Numerical Modelling of One-Dimensional Laminar Pulverized ... .fr

heat flux between particles; Heff is the effective heat of reaction; Qw is the radiative heat ...... 405-422. [84] Solomon, R.P., Serio, M.A., and Suuberg, E.M. (1992).
410KB taille 2 téléchargements 383 vues
Numerical Modelling of One-Dimensional Laminar Pulverized Coal Combustion Phần II (1995)

Trịnh Minh Chính

1

1. INTRODUCTION Research into coal dust combustion has been going on for almost a century. Everybody agrees that coal particles burn in a very complicated manner. The combustion process includes

devolatilization

(pyrolysis),

homogeneous

reactions

(reactions

of

volatile

components), heterogeneous reactions (surface reactions or char oxidation), swelling, cracking, etc. Many research workers have attempted to simplify the process by deviding it into several subprocesses, e.g. devolatilization, char oxidation, etc.. Kobayashi et al. (1976) measured weight loss of lignite and bituminous coal by devolatilization at high temperatures. They proposed a pyrolysis model which consists of two competing overall reactions instead of a single first-order reaction. According to Pitt (1962), any theory of the kinetics of coal pyrolysis must be based on a number of simplifying assumptions, namely that coal is a mixture of many components which decompose independently, and that the composition reactions have a wide range of activation energies. He developed a new model for coal devolatilization called Distributed Energy Activation Model (DAEM), based on total volatiles. Anthony et al. (1974) correlated their experimental data of lignite and bituminous coals to find parameters for this model. Solomon et al. (1988) suggested a model applicable for all volatile components, e.g. CO, H2O, CO, HCN, C2H6, H2, etc. There are many models of coal devolatilization such as multiple parallel reaction model (MPRM), distributed activation energy model (DAEM), multiple competing reaction model (MCRM), consecutive competing char-forming reactions model, etc. Reviews of devolatilization of coal particles are described in Saxena (1990), Solomon et al. (1992), Essenhigh (1981) and Anthony and Howard (1976). The combustion of volatile components is very complicated. The DAEM and MPRM may give more details of volatile components compared with the other models. However, the composition of volatiles (mass fractions of single volatile or volatile components) for different types of coal are usually unknown and they must be calculated. Some simple methods have been mentioned in Gray et al. (1976). The composition of volatiles in this report was calculated by an interpolated method (APP.A1). It is based on the experimental data of Solomon et al. (1992). Reactivity and combustion of coal chars is also a complicated subject. The burning rate of coal char depends on the rate of transport of oxygen to the external surface of the particle to react with fixed carbon and form both CO and CO2. CO begins to burn if the temperature is greater than about 923 K, and CO is the sole heterogeneous product when the particle temperature is above 1373 K. According to Saxena (1990), two important questions in establishing the combustion mechanism for a given combustion system are first, what are the combustion products on the char particle's surface? and second, which of the two gases, oxygen or carbon dioxide, should preferably diffuse to the surface? From this information, several models, e.g. one-film, two-film and continuous-film are proposed, depending on the behaviour of CO and CO2 at the boundary layers surrounding small particles. According to Mitchell et al. (1990), no definitive experimental data exist at high temperatures to establish

2

the relative importance of CO2 formation on the surfaces of particles and CO2 formation in the boundary layers surrounding particles in the pulverized-coal size range. The term "modelling", according to Spalding (1963), is used to connote the practice of predicting the likely results of one experiment by way of the interpretation of the results of another experiment. Modelling of coal combustion, including pyrolysis, combustion of volatiles, char oxidation, soot formation and burning, mass transfer, heat transfer, etc., is a very complicated matter. Several models for pyrolysis, char oxidation, and a number of reactions for volatile oxidation have been proposed. Each model has several empirical constants which are found under certain conditions. These constants or models must be tested in different cases with different conditions. Smoot et al. (1976a) have studied the propagation of laminar, coal-dust air flames. They predicted flame velocity, temperature, species concentration (CO, CO2, H2O, H2, H, OH, O, O2) and particle diameter. Effects of gaseous diffusion, coal pyrolysis, char oxidation, and gasous reactions were considered. Smith and Smoot (1980) have developed a onedimensional model for pulverized coal combustion and gasification processes. The heat transfer between gas and particles, particles and particles, devolatilization and char oxidation were described very thoroughly. Smoot (1984,1991) made a review of models for pulverized coal combustors/gasifiers, one-dimensional and multidimensional coal reactors, fixed bed and one-dimensional fludized bed. In this report, both Illinois #6 and Pittsburgh #8 coals were tested. Two devolatilization models called multiple parallel reaction model (MPRM) and distributed activation energy model (DAEM) have been used for the investigation. The char oxidation model is described by a single-film model, i.e. both CO and CO2 are the heterogeneous products and CO reacts with oxygen after it has been transported to the free stream. Two different "reaction rates" were tested. The first one called char oxidation model 1 or Mitchell's model is described in Mitchell (1988) and the second, simple model called char oxidation model 2, is described in Field et al. (1967). It is called a simple model because the apparent reaction order equals unity. Six sets of kinetic reaction model are described for combustion (oxidation) of active volatile components (CO, C2H6, CH4, H2 and HCN). They are called global model, reduced model 1, 2, 3 and 4, and elementary model. They varied from 3 reactions (global model) to 123 reactions (elementary model). Except for the elementary model, all kinetic reaction models were tested with and without the simple char oxidation model. The elementary model with 123 reactions is too comprehensive to be calculated with the char oxidation model. Different oxygen concentrations in oxidizer and different particle sizes are investigated. The influences of radiation of particles to a cold environment and radiation between particles are also considered. The governing equations, devolatilization models, and char oxidation models are described in detail. APP.A1 shows the method for calculating the composition of volatiles. The numerical calculation of DAEM for the non-isothermal process is described in APP.A2. The solver method (Newton-Rapshon method) for oxidation model 1 (Mitchel's model) is

3

described in the subsection "Combustion of Coal Char". Results from the calculations were compared with experimental data. These comparisons are shown in concentration and temperature profile figures. Some of the particle profiles obtained with different methods are also compared with each other as well as with experimental data, and are shown in the section "Results and discussions". APP.B1 shows the results of the calculations for Illinois #6 coal air flame. APP.B2 shows the results of the calculations for Pittsburgh #8 coal air flame. APP.C shows the figures of the proximate and ultimate analysis (dry-ash-free basic) of coal char, weight loss of chemical elements of coal char as a function of particle mass loss, and particle mass and mass loss of coal particles in coal air flames. These figures are purely illustrative and therefore need no comment. APP.D shows the analysis of particle sizes for six coal types: Blair Athol, Cerrejon, Illinois #6, Middleburgh, Pittsburgh #8 and Ulan.

4

2 MATHEMATICAL MODEL In order to overcome the complexity of pulverized-coal combustion, including fluid mechanics, devolatilization, heterogeneous (particles and gas) and homogeneous (gas) reactions, a model is developed with the following assumptions: 1- The flow is one-dimensional, steady, and premixed. 2- The measured gas temperatures are used to compute reaction rates. The gas temperatures

are

known

(measured)

so

that

they

cannot

influence

the

kinetic

(homogeneous) reactions. 3- The velocities of gas and particles are equal. 4- Body forces and viscous forces within the gas phase are disregarded. 5- Coal particles are treated as spheres. 6- Tar and soot formations are not considered. 7- Swelling formation is not considered. 8- The particle size is constant and equals the mass-mean particle size. 9- The particle temperature is uniform. 10- Oxygen is transported to the surface and products are transported away by molecular diffusion. 11- Molecular diffusion in the gas phase is not influenced by the presence of coal particles.

2.1 Governing equations The governing equations for one-dimensional laminar coal combustion (Fig. 2.1) can be expressed as follows.

2.1.1 Particle phase Mass balance equation

L d&M pj Atot ⎧ ⎫ + ⎨ n j A pj q + M& pj,0 ∑ Y i ⎬ = 0 i=1 ⎭ dx V tot ⎩

(1)

Energy balance equation

⎛ dq+ dq - ⎞ B dT p Atot ⎟ & ( ) + n h C Atot ⎜⎜ j A pj g Tg Tp M pj pj ⎟ 1 - eB dx V tot ⎝ dx dx ⎠

- Atot n j A pj q H eff + Atot Q w = 0 V tot V tot

(2)

where x is the spatial coordinate; Mpj is the mass flow rate of particles; Atot is the cross3

sectional area; Vtot (= Atot dx) is the volume; nj is the number of particles per cm ; q is the overall particle burning rate (the rate of consumption of carbon); Mpj,0 is the initial particle mass; Yi is the mass fraction of species i, released by pyrolysis; Tp is the particle temperature; Cpj is the constant pressure heat capacity of particles; Apj is the particle surface area; hg is the heat transfer coefficient; Tg is the gas temperature; qr is the radiative heat flux between particles; Heff is the effective heat of reaction; Qw is the radiative heat transfer from hot particles to the cold environment.

5

Figure 2.1: Flat-flame burner schematic The first term on the left side of Eq.(1) is the mass rate of change of particles. The second term is particle mass loss due to char burning. The last term is particle mass loss due to pyrolysis. The first term on the left side of Eq.(2) is the change of particle temperature. The second term, the convection term, is the heat transfer between particles and gas. The third term is the radiative heat transfer between particles. The next term is the production (source) term, the energy being generated by the burning of char. The last term is the energy, which is lost in a cold environment. The effective heat of reaction Heff can be calculated from the heat of combustion of the two reactions, C + ½O2 = CO and C + O2 = CO2.

H eff = ψ • H CO2 + (1 - ψ ) • H CO

(3)

where ψ and (1-ψ) are the fractions of the carbon content of the particle to be converted to CO2 and CO at the particle surface. HCO and HCO2 are the heat of reaction per mole carbon 10

for the reactions C + ½O2 = CO and C + O2 = CO2, respectively. HCO = 9.79x10 10

(2340.0 cal/g) and HCO2 = 33.05x10

ergs/g

ergs/g (7900.0 cal/g) (Ayling and Smith (1972).

The heat transfer coefficient is defined as

6

hg =

Nu λ g

(4)

dp

where Nu is the Nusselt number; λg is the thermal conductivity of the mixture, and dp is the particle diameter. For a spherical particle in the gas stream, Nu is found as (Ranz and Marshall (1952)) 1

1

Nu = 2.0 + 0.6 ReD2 Pr 3

(5)

Nu is given as 2 since particle and gas have the same velocities. The radiative influence on particles by the cold environment is

Q w = n j A pj ε σ ( T 4p - T 4w )

(6)

ε is the particle emissivity; σ is the Stefan-Boltzmann constant; Tw is the temperature of the room wall. By considering the influence of mass transfer on heat transfer using film theory, the Stefan flow parameter B is found as (Water et al. (1988a,a988b))

B= -

(1 -ν ) q d p C pg 2 Mcλ

(7)

where v is the oxygen stoichiometric coefficient. For small particles, B can be taken as 1. The specific heat at constant pressure (constant pressure heat capacity) of the particles in the flame is calculated by JANAF Thermochemical Tables For Tp = 300 - 1000K:

2

Cpj/R = - 0.6705661 + 0.07181499E-1⋅T - 0.05632921E-4⋅T 3

4

+ 0.02142298E-7⋅T - 0.04168562E-11⋅T For Tp = 1000 - 5000K:

2

Cpj/R = 1.49016640 + 0.16621256E-2⋅T - 0.06687204E-5⋅T 3

4

+ 0.12908796E-9⋅T - 0.09205334E-13⋅T

7

Figure 2.2: Specific heat of particles in Illinois #6 coal air flame. Figure 2.2 shows the calculated specific heat of the carbon in Illinois #6 coal air flame. The specific heat of particles varies from 0.7 for Tp = 298K to 2.1 kJ/(kg K) for Tp = 1771K. It is clear that the specific heat of particles cannot be taken as a constant. The particle temperatures calculated with constant and varied Cpj are shown in Fig.3.3 (see section 3 Results and Discussions). The difference particle temperature calculated with two different Cpj is about 50.0 K (near the peak of the temperature profile).

2.1.2 Radiative model

Figure 2.3: Schematic of two-flux model The two-flux model of radiative heat transfer between particles can be written as (Siddal (1974))

8

+

dq = - 2 ( σ a + b σ s ) q+ + 2 σ a σ T 4p + 2b σ s qdx dq = - 2 ( σ a + b σ s ) q- + 2 σ a σ T 4p + 2b σ s q+ dx +

q

(8) (9)

-

and q are the radiative heat fluxes in positive and negative directions, respectively +

(Fig.2.3). On the left-hand side of Eq.(8) is the change of q . The first term on the right+

hand side is the extinction of q due to absorption and back-scattering. The next term is the source term due to hot particles which emit radiation. The last term is the back scattering of -

+

q contributing to the q . The interpretation of Eq.(9) is similar to that for Eq.(8), but in the opposite direction. Combining Eqs.(8) and (9) yields +

-

dq dq = - 2 σ a ( q+ + q- ) + 4 σ a σ T p4 dx dx where

σ a = Ca σ s = Cs

π 4

π

4

(10)

2

(11)

2

(12)

d p nj d p nj

σa and σs are the absorption and scattering coefficients for monodisperse coal dust cloud, respectively; Ca is the absorption constant; Cs is the scattering constant.

2.1.3 Gas phase The mass balance equation

L d&M g Atot ⎧ ⎫ ⎨ n j A pj q + M& pj,0 ∑ Y l ⎬ = 0 l=1 ⎭ dx V tot ⎩

(13)

The energy balance equation (this equation is not used in the calculations)

1 d ⎧ dT g dT g ⎫ dT g Atot K ρ ∑ Y k V k C pk ⎨ λ Atot ⎬+ M& g dx dx ⎭ dx C p dx ⎩ C p k=1 K 1 B A + tot ∑ ω& k hk W k + Atot hg ( T g - T p ) = 0 n j A pj 1-e B V tot C pg C pg k=1

(14)

Species conservation

dY k d { + ρ Atot Y k V k} - Aω& k W k = 0 M& g dx dx

(15)

where

ρ=

pW RT

(16)

Mg is the mass flow rate of gas; Cp is the constant pressure heat capacity of the mixture; Yk is the mass fraction of the kth species (there are K species); Cpk is the constant pressure

9

heat capacity of the kth species per unit volume; ω& is the molar rate of production by chemical reaction of the kth species per unit volume; hk is the specific enthalpy of the kth species; Wk is the molecular weight of the kth species; W is the mean molecular weight of the mixture; R is the universal gas constant; Vk is the diffusion velocity of the kth species. Fig.2.4 shows the calculated gas mass density in Illinois #6 coal air flame. The gas mass 3

density decreases from 0.001185 g/cm at x = -1 cm (distance before the burner outlet, Tg 3

= 298 K) to 0.000898 g/cm at x = 0 cm (at the burner outlet, Tg = 410 K) and 0.000180 3

3

g/cm at x = 0.5 cm (Tg = 1770 K), and it increases to 0.000232 g/cm at x = 3 cm (Tg = 1327 K). Table 2.1 shows the gas mass density in coal air flame and the air density for the same gas temperature. The difference between them is about 10% at the high temperature (T = 1770.0 and 1340.0 K). At the low temperature (T = 298.0 and 410.0 K), the gas density and air density have the same values due to the air being present only in the gas phase of the flame. Tg (K)

298.0

410.0

1770.0

1340.0

X (cm)

-1.0

0.0

0.5

3.0

Gas mass density

0.001185

0.000852

3

(g/cm )

0.000232 0.000180

3

Air density (g/cm )

0.001185

0.000852

0.000260 0.000197

Table 2.1: Comparison between gas density in coal air flame and air density.

Figure 2.4: The calculated gas mass density in Illinois #6 coal air flame. The diffusion velocity, Vk, may be composed of three parts: the ordinary diffusion velocity, Vkd, the non-zero thermal diffusion velocity, Vkt, and the correction velocity, Vc

10

V k = V kd + V kt + V c

(17)

Vkd is given by Curtiss and Hirschfelder (1949) as

D k dX k X k dx

V kd = -

(18)

where Xk is the mole fraction, and where the mixture-averaged diffusion coefficient Dk is given explicitly in terms of the binary diffusion coefficient Djk.

1-Yk K X j ∑ j ≠ k D kj

Dk =

(19)

The non-zero thermal diffusion velocity Vkt is given

V kT =

D k k Tk 1 dT X k T dx

(20)

where kkT is the thermal diffusion ratio (Chapman and Cowling (1970)). The correction velocity Vc (independent of species but a function of x) is included to ensure that the mass fractions toal unitty. More details of diffusion velocities can be seen in Kee et al. (1985,1992).

2.1.4 Chemical reaction rate expressions (Kee et al. (1985)) Consider i elementary reversible (or irreversible) reactions involving K chemical species that can be represented in the general form K

K

k =1

k =1

∑ν k ′i X k = ∑ν k ′′i X k

(i = 1,..., I)

(21)

the stoichiometric coefficient vki are integers and χk is the chemical symbol for the kth species. The production rate ωk of the kth species can be written as a summation of the rate-ofprogress variables for all reactions involving the kth species: I

ω& k = ∑ ν ki qi i=1

11

(k = 1,..., K)

(22)

where

ν ki = (ν k ′′i - ν k ′i )

(23)

The rate-of progress variable qi for the ith reaction is given by the difference of the forward rates and the reverse rate as K

K

k =1

k =1

qi = k f i Π [ X k ]ν k′i - k ri Π [ X k ]ν k′′i

(24)

where [Xk] is the molar concentration of the kth species and kfi and kri are the forward and reverse rate constants of the ith reaction. The forward rate constants for the I reactions are generally assumed to have the following Arrhenius temperature dependence:

⎧ Ei ⎫ β k f i = Ai T exp⎨⎬ ⎩ RT ⎭

(25)

where Ai is the pre-exponential factor, βi is the temperature exponent, and Ei is the activation energy. The reverse rate constants kri are related to the forward rate constants through the equilibrium constant as

k ri =

k fi

(26)

K ci

Although Kci is given in concentration units, the equilibrium constants are more easily determined from the thermodynamic properties in pressure units; they are related by K

ν ki ⎛ P atm ⎞k∑ ⎟ =1 K ci = K pi ⎜ ⎝ RT ⎠

(27)

where Patm denotes a pressure of 1 atm. The equilibrium constants Kpi are obtained with the relationship

⎧ ∆ S io ∆ H io ⎫ K pi = exp⎨ ⎬ RT ⎭ ⎩ R

(28)

the ∆ refers to the change that occurs in passing completely from reactants to products in the ith reaction. More specifically,

12

o K ∆ S io S = ∑ ν ki k k =i R R o o K ∆ Hi = ∑ ν ki H k k =i RT RT

(29) (30)

2.1.5 Discretization (Kee et al. (1985)) All partial differential equations are rewritten to a system of algebraic equations by using the finite difference approximation. The initial approximations are usually on a very coarse mesh that may have few points. After obtaining a solution on the coarse mesh, new mesh points are added in regions where the solution or its gradients change rapidly. An initial guess for the solution on the finer mesh is obtained by interpolating the coarse mesh solution. This procedure continues until no mesh points are needed to resolve the solution to the degree specified by the user. The system of algebraic equations is solved by the damped modified Newton algorithm. However, if the Newton algorithm fails to converge, the solution estimate is conditioned by a time integration. This provides a new starting point for the Newton algorithm that is closer to the solution, and thus more likely to be in the domain of convergence for Newton's method. The upwind (windward) difference scheme is used on the convective terms, while the second-order centered differentiation is used on the diffusive term. The diffusive term in the species conservation equation is approximated in a similar way, but it appears to be different because of the diffusion velocities. More details concerning discretization can be found in Kee et al. (1985).

2.1.6 Boundary conditions At the cold boundary (inlet), the mass flux fractions and the temperature can be written

Y k,1 + ( ρA Y k V k ) j=1 1 = 0

(31)

T1 -Tb = 0

(32)

2

and

where Tb is the specified burner temperature. At the hot boundary (outlet), all gradients vanish, i.e.

Y k, j - Y k, j -1 =0 x j - x j -1 T k, j - T k, j -1 =0 x j - x j -1

(33) (34)

At the cold and hot boundaries, the radiative heat fluxes in positive and negative directions are

13

x < 0 : q+ = 0 ; q- = 0 x≥0: q = ε σ T

4 p

x≥0: q = q

; q =σ T

+ 0

+ J

+

+ J -1

0

; q = q J

-

(35) 1

(36)

4 p

(37) -

where q0 and q0 are the flux densities at the burner outlet; q1 is the flux density at the first node point after the burner outlet; qJ is the flux at the end of the flame; Tp is the particle temperature.

2.2 Coal devolatilization (pyrolysis) Coal devolatilization is important in a coal conversion process. The volatiles, which can account for up to 70% of the coal's weight loss, control the ignition, the temperature and the stability of the flame in combustion (Water et al. (1988a)). According to Field et al. (1967), when coal is heated to a sufficiently high temperature it begin to decompose, producing tars and gases and volatiles. The volatiles consist of a mixture of combustible gases, carbon dioxide, and water vapour. Apart from carbon monoxide and hydrogen the combustible gases are mainly hydrocarbons, although there are small quantities of phenolic and other compounds. The volatile yield and composition are influenced by the coal rank and type, and for example, by the externally specified variables such as particle size, heating rate, time-temperature history and pressure (Carpenter and Skorupska (1993)). o

The decomposition process with the heating rates less than or equal 10 C/min is called slow decomposition (used in industrial carbonization processes) and the decomposition with the heating rates higher than those of slow decomposition is called rapid decomposition (as happens in most combustion systems, e.g. fluidized bed, fixed bed, pulverized coal furnaces, etc.) (Field et al. (1967)). Many experiments have been made to study the kinetics of coal pyrolysis, e.g., 1) Heated grid experiments (Anthony et al. (1974), Suuberg et al. (1979)); 2) Entrained flow reactors: Drop tube furnace (Maloney and Jenskins (1984), Solomon et al. (1982)); Heated tube reactor (Serio et al. (1987)); Transparent wall reactor (Field (1970), Fletcher (1989b)); Well-stirred reactors (Goldberg and Essenhigh (1989)); 3) Thermogravimetric analysis (TGA) (Carangelo et al. (1987), Ottaway (1982), Serio et al. (1987)); 4) Radiative heating (Hertzberg and Zlochower (1991)); 5) Fluidized bed (LaNauze (1982), Pitt (1962)); 6) Fixed bed (Gokhale et al. (1986), Parikh and Mahalingam (1987). Rewiews of coal devolatilization can be seen in (Anthony and Howard (1976), Essenhigh (1981), Howard (1981), Solomon et al. (1992)). According to Anthony and Howard (1976), several authors at different times have reviewed many of the diverse findings, each time identifying new aspects of the problem and new directions for research. The recent acceleration of coal research has generated much new information which must be reconciled and reviewed. A number of models of coal pyrolysis have been proposed, e.g., a) Single Reaction: nth-Order Arrhenius reaction, Non-

14

Arrhenius model; b) Multiple Parallel Reaction: Two first-order Arrhenius reactions, Multiple first-order reactions with statistical distribution of activation energies; c) Multiple Competing Reactions: Two-first-order reactions, Multiple first-order reaction; d) Complex Schemes: Multiple consecutive parallel first-order reactions, Parallel competing first-order reactions; e) Schemes involving secondary char-forming reactions: Consecutive competing charforming reactions, Parallel competing char-forming reactions. More details of these models can be found in Essenhigh (1981). Here is a brief description of the multiple parallel reaction model (MPRM) and distributed activation energy model (DAEM), which are used in the calculations.

2.2.1 Multiple Parallel Reaction Model (MPRM)(Anthony et al. (1974) a) Model based on total volatiles The simplest model for coal devolatilization based on the quantity of volatiles (total weight loss by the particles) can be expressed as

dV = k ( V * - V) dt

(38)

*

where V is the mass of volatiles, per mass of original coal, evolved at time t, V is the value of V at t = ∞ and k is the rate constant. It may be written

k = k 0 exp{-

E } RT

(39)

where k0 is the apparent frequency factor, E is the activation energy, R is the ideal gas 16

constant, and T is the absolute temperature. k0 = 6.5x10

-1

sec , E = 36.89 kcal/mole

(Anthony et al. (1974). For a given set of conditions, values of k from various authors may differ by several factors of 10 and E may vary from several kcal/mole to nearly 50 kcal/mole. b) Model based on individual volatile constituents Pitt (1962) assumed that coal decomposes thermally as if it were a mixture of many pseudospecies, each of which decomposes via an independent first-order reaction and with a characteristic energy, i.e. (Saxena (1990))

dV i = k i ( V *i - V i ) dt

(40)

Ei } RT

(41)

where

k i = k 0i exp{-

subscript i denotes the composition of the volatile matter, e.g. CO2, CO, H2O, etc. The rate

15

constants for MPRM are shown in Table 2.2.

Species

E0

k0 -1

[sec ]

Ref.

[kcal/mol] 16

67.0

[69,84]

18

60.0

[69,84]

18

79.0

[69,84]

13

59.6

[69,84]

14

59.1

[69,84]

13

59.0

[69,84]

14

80.0

[69,83,84]

12

13.7

[69,83,84]

CO2

6.5 x 10

CO

2.2 x 10

H2O

1.4 x 10

HCN

1.7 x 10

C2H6

8.4 x 10

CH4

7.5 x 10

H2

1.0 x 10

NH3

1.2 x 10

H2S

2.91 x 10

Tar

4.3 x 10

9

36.89

[1]

14

54.6

[69]

14

27.5

[84]

8.6 x 10

Table 2.2: The rate constants of MPRM based on individual volatile constituents. These constants are taken from DAEM with the standard deviations, σi, are set zeroes.

2.2.2 Distributed Activation Energy Model (DAEM) (Anthony et al. (1974), Saxena (1990), Serio et al. (1987), Solomon et al. (1988,1992)) a) Model based on total volatiles DAEM can also be called multiple first-order reactions with statistical distribution of activation energies. The activation energy, E, in Eq.(39) is expressed as a continuous distribution function f(E) so that f(E)⋅dE is the fraction of potential volatiles having an activation energy between E and E + dE. Hence for all volatiles * * dV = V f(E) dE

with

(42)



∫ f(E) dE = 1

(43)

0

Anthony et al. (1974) proposed a Gaussian distribution for the activation density function f(E) so that

16

f(E) =

1 ⎧ (E - E 0 ) ⎫ exp⎨‰ 2 ⎬ σ (2π ) ⎩ 2σ ⎭

(44)

Combining Eqs.(42), (43) and (44) yields * V -V ∞ ⎧ ∞ ⎫ = exp ∫ ⎨ - ∫ k t dt ⎬ f(E) dE * 0 ⎩ 0 ⎭ V

(45)

For simplicity, the integral limits were changed to E0 ± 2σ which includes 95.5% of the reaction set. Eq.(45) then becomes * V - V E max ⎧ ∞ ⎫ = exp ∫ ⎨ - ∫ k t dt ⎬ f(E) dE * ⎩ 0 ⎭ V E min

(46)

E0 is the mean activation energy. σ is the standard deviation. Emin = E0 - 2σ and Emax = E0 + 2σ. 16

where k0 = 6.5x10

-1

sec , E = 36.89 kcal/mole and σ = 4.18 kcal/mole (Anthony et al.

(1974)). For the isothermal process, Eqs.(45) and (46) become * V -V ∞ = ∫ exp{- k t} f(E) dE * 0 V

(47)

* V - V E max = ∫ exp{- k t} f(E) dE * V E min

(48)

b) Model based on individual volatile constituents The DAEM for the nonisothermal process based on individual volatile constituents can be written

V i -V i ∞ ⎧ ∞ ⎫ = exp ∫ ⎨ - ∫ k i t dt ⎬ f( E i ) dE * 0 ⎩ 0 ⎭ Vi

(49)

V i - V i Ei,max ⎧ ∞ ⎫ = exp ∫ ⎨ - ∫ k i t dt ⎬ f( E i ) dE * ⎩ 0 ⎭ Vi E i,min

(50)

*

or *

where

k i = k 0i exp{-

17

Ei } RT

(51)

f( E i ) =

⎧ ( - )⎫ 1 exp⎨- E i E2 0 ⎬ ‰ σ i (2π ) ⎩ 2σ i ⎭

(52)

For the isothermal process, Eqs.(49) and (50) become

V i -V i ∞ = ∫ exp{- k i t} f( E i ) dE * 0 Vi *

(53)

*

V i - V i Ei,max = ∫ exp{- k i t} f( E i ) dE * Vi E i,min

(54)

The rate constants for this model are shown in Table 2.3

2.2.3 Q-factor The yield of volatiles in the models was greater than the potential yield indicated by the proximate volatile matter of the coal. Jensen and Mitchell (1993) defined the Q-factor as the ratio of actual volatile yield to the proximate volatile matter. They measured the volatiles release in 6 and 12 mole-% oxygen environments for 9 different coal types (Table 2.4 (dry basic) and Table 2.5 (dry-ash-free basic)). More details are given in (Essenhigh (1981)).

Q - factor =

Species

E0

k0 -1

[sec ] CO2 CO H2O HCN C2H6 CH4 H2 NH3 H2S Tar

Actual volatile yield Proximate volatile matter σ

[kcal/mol]

(55)

Ref.

[kcal/mol]

16

67.0

3.0

[69,84]

18

60.0

3.0

[69,84]

18

79.0

11.8

[69,84]

13

59.6

3.0

[69,84]

14

59.1

3.0

[69,84]

13

59.0

3.9

[69,84]

14

80.0

12.5

[69,83,84]

12

13.7

1.5

[69,83,84]

6.5 x 10

2.2 x 10 1.4 x 10

1.7 x 10 8.4 x 10

7.5 x 10 1.0 x 10

1.2 x 10

9

2.91 x 10

36.89

4.18

14

54.6

0.0

[69]

14

27.5

1.5

[84]

4.3 x 10 8.6 x 10

[1]

Table 2.3: The rate constants of MPRM based on individual volatile constituents (Solomon et al. (1992)).

18

Q-factor

ASTM

(dry basic)

Volatile

Measured volatile mass loss

% (dry Coal type

basic) 6 mole% O2 Volatile mass loss % (dry)

12 mole% O2 Q-

Volatile mass

factor loss % (dry)

Qfactor

Cerrejon

36.8

55

1.49

53

1.44

Blair Atholl

28.6

38

1.33

41

1.43

Ulan

31.0

42

1.35

42

1.35

PSOC 1445d Blue #1

43.5

51

1.17

51

1.17

PSOC 1451d Pittsburgh #8

32.6

49

1.50

49

1.50

PSOC 1488d Dietz

45.1

54

1.20

57

1.26

PSOC 1493d Illinois #6

36.4

51

1.40

50

1.37

PSOC 1502d Hiawatha

38.8

54

1.39

55

1.42

PSOC 1516d L.Kittaning

18.1

13

0.72

11

0.61

Table 2.4: Measured volatile mass loss and Q-factor in 6 and 12 mole-% oxygen environments (dry basic) (Jensen and Mitchell (1993)). Q-factor (daf basic)

ASTM

Measured volatile mass loss

Volatile Coal type

% (daf basic) 6 mole% O2

12 mole% O2

Volatile mass

Q-

loss

factor

Volatile mass

Q-

loss

factor

% (daf)

% (daf)

Cerrejon

41.3

60.9

1.47

58.7

1.42

Blair Atholl

31.4

42.2

1.34

45.6

1.45

Ulan

37.3

48.9

1.31

48.9

1.31

PSOC 1445d Blue #1

45.7

52.9

1.16

52.8

1.16

PSOC 1451d Pittsburgh #8

38.4

54.6

1.42

55.4

1.44

PSOC 1488d Dietz

47.6

56.4

1.18

59.6

1.25

PSOC 1493d Illinois #6

43.1

57.4

1.33

56.1

1.30

PSOC 1502d Hiawatha

42.3

58.1

1.37

59.0

1.40

PSOC 1516d L.Kittaning

20.2

18.3

0.91

15.5

0.77

Table 2.5: Measured volatile mass loss and Q-factor in 6 and 12 mole-% oxygen environments (dry-ash-free (daf) basic) (Jensen and Mitchell (1993)).

2.3 Combustion of volatiles 19

Although the combustion properties of fuels have been studied empirically for many years, early numerical combustion modeling virtually ignored chemistry. In the last decade, numerical modelling has rapidly become an essential part of many combustion research and development programme, and there has been an accelerating evolution from the use of single-step empirical representations, to the use of lumped (overall) multistep models, and finally to the inclusion of full detailed chemical kinetic mechanisms to better simulate chemistry interactions (Dryer (1991), Westbrook and Dryer (1984)). For two- or threedimensional geometry and turbulent combustion systems, full detailed chemical kinetic mechanisms with many reactions cannot currently be used because of the computational costs. A detailed chemical kinetic mechanism with few reactions (which some authors call quasi-global models) may be a best solution for accuracy and computational costs. Some global models and multistep reaction (quasi-global) models have been investigated and examined. More details of chemical kinetic mechanism in combustion systems can be found in Basevich (1987), Baulch et al. (1994), Glassman (1977), Warnatz (1984), Westbrokk and Dryer (1981b), Westbrook and Pitz (1984).

2.3.1 Composition of volatiles The coal particles are assumed to be surrounded by air (premixed) and the volatiles mix with the air as they are released from the particles. If the gas temperature is sufficiently high, ignition will be spontaneous. A knowledge of the composition (mass fraction) of volatiles which evolve from coal particles is very important for the calculation of volatiles. Suuberg et al. (1979) have measured the composition of volatiles and of the formation kinetics of each product released from a lignite and a bituminous coal. They indicated that the lignite volatiles are dominated by CO, CO2 and H2O, while the main volatiles from the bituminous coal are tar and light hydrocarbons. In general, pyrolysis water, carbon dioxide and tar evolve at low temperatures, whereas hydrocarbons, CO and hydrogen evolve at higher temperatures. In bituminous coal, tar constitutes 50-80% of the released volatiles, the remaining volatiles consisting of hydrocarbon gases, water and oxides of carbon (Saxena (1990)). Several other investigations of the composition of volatiles have been made, e.g. Loison and Chauvin (1964), Serio et al. (1987). A review of investigations on product distribution is mentioned in Howard (1981). According to Solomon et al. (1992), carbon dioxide (CO2), moisture (H2O), carbon monoxide (CO), hydrogen cyanide (HCN), ethylene (C2H6), methane (CH4), and hydrogen (H2) are the major gases (volatiles) from Illinois #6 coal. There are five active gases, CO, C2H6, CH4, H2 and HCN, in volatiles, three of which (CO, CH4 and H2) are most significant for coal combustion due to their high activities. Carbon monoxide is a most important gas, resulting partly from the pyrolysis, partly from the char oxidation. The following reactions and their rate constants may be used for calculating the combustion of volatiles. More details of these reactions are given in the relevant references. The original references of the multistep reactions are given in Miller et al. (1982), Westbrook and Dryer (1979), Westbrokk and Pitz (1984).

2.3.2 Global models

20

a) Oxidation of CH4 The oxidation of a single species, e.g. CO, CH4, H2 etc., expressed by a single reaction is called global reaction. Many global reaction models for oxidations of CH4 and CO have been studied. The global reaction for oxidation of CH4 can be written as

CH 4 + 2 O 2 → CO 2 + 2 H 2 O

(56)

The overall methane disappearance-rate expression in a turbulent-flow reactor at atmostpheric pressure, over the temperature range 1100 - 1400 K, found by Dryer and Glassman (1973) is

⎧ 48400 ⎫ d[ CH 4 ] 0.70 0.80 = - 1.58x 1013 exp⎨⎬ [C H 4 ] [ O 2 ] dt ⎩ RTg ⎭

(57)

Westbrook and Dryer (1981a) have developed one-step and two-step reaction mechanisms for hydrocarbon fuel (CnH2n+2, n = 1-10) using a numerical laminar flame model. They proposed a one-step reaction mechanism representing the oxidation of a conventional hydrocarbon fuel

and its rate constant

Fuel + n1 O 2 → n2 CO 2 + n3 H 2 O

(58)

⎧ Ea ⎫ a b n k ov = A T exp⎨⎬ [Fuel ] [Oxidizer ] ⎩ RT ⎭

(59)

The one-step reaction mechanism for hydrocarbon oxidation in a turbulent flow reactor can also be written (Westbrook and Dryer (1981a))

m⎞ m ⎛ C n H m + ⎜ n + ⎟ O 2 → n CO 2 + H 2 O 4⎠ 2 ⎝

(60)

and a two-step reaction mechanism

m ⎛n m⎞ C n H m + ⎜ + ⎟ O 2 → nCO + H 2 O 2 ⎝2 4 ⎠ 1 CO + O 2 = CO 2 2

(61) (62)

The rate of the CO oxidation reaction is taken from Dryer and Glassman (1973). In order to reproduce both the proper heat of reaction and pressure dependence of the [CO]/[CO2] equilibrium, a reverse reaction was defined for Eq.(62), with a rate 1.0 8 k CO2→CO+‰O2 = 5 • 10 exp(-40,000/RT) [ CO 2 ]

21

(63)

For methane air flame (n = 1, m = 4), Eq.(61) becomes Eq.(56). The rate constants for methane air flame are shown in Table 2.6.

Methane flame

Ea

a

b

[kcal/mole]

[-]

[-]

8

48.4

-0.3

1.3

5

30.0

-0.3

1.3

9

48.4

-0.3

1.3

7

30.0

-0.3

1.3

A 3

[mole/(cm s)] One-step

1.3x10 8.3x10

Two-step

2.8x10 1.5x10

Table 2.6: The rate constants for methane air flame (Westbrook and Dryer (1981a)). According to Westbrook and Dryer (1981a), the two-step mechanism predicts flame speeds of methane-air (hydrocarbon-air) flames in close agreement with those predicted by the one-step model. However, the negative exponents can give problems in computation. Williams et al. (1969) have studied the combustion of premixed methane-oxygennitrogen (CH4-O2-N2) mixtures in the range T = 1400-1800 K and P = 0.3-1.1 atm in a 3

small (2.9 cm ) conical reactor fed by a single-choked sonic-flow jet. The burning-rate for one-step reaction is

⎧ 57000 ⎫ d[ CH 4 ] 0.5 0.5 = 5.3x 1012 exp⎨⎬ [ CH 4 ] [ O 2 ] [ H 2 O ] dt ⎩ RT g ⎭

(64)

and for two-step reaction

⎧ 60000 ⎫ d[ CH 4 ] 0.5 0.5 = 1.7x 1013 exp⎨⎬ [ CH 4 ] [ O 2 ] [ H 2 O ] dt ⎩ RT g ⎭

(65)

⎧ 20000 ⎫ d[CO] 0.5 0.5 = 3.5x 106 exp⎨⎬ [CO] [ O 2 ] [ H 2 O ] dt ⎩ RT g ⎭

(66)

b) Oxidation of CO The global reaction for CO oxidation can be written

1 CO + O 2 → CO 2 2

(67)

Dryer and Glassman (1973) have studied carbon monoxide oxidation in the presence of water at atmospheric pressure, over the temperature range 1030 - 1230 K, and over water concentrations of 0.1% - 3.0%. With the linear least-square-fitting techniques, the rate

22

constant is found as

⎧ 40000 ⎫ d[CO] 0.25 0.5 = 3.98x 1014 exp⎨⎬ [CO] [ O 2 ] [ H 2 O ] dt ⎩ RT g ⎭

(68)

Howard et al. (1973) have developed a global model for CO oxidation from their flat methane-air flame experimental results and the previous data (Eq.(14)). The model is adequately correlated for the flame with the temperature range 840 - 2360 K.

⎧ 30000 ⎫ d[CO] 0.5 0.5 = 1.3x 1014 exp⎨⎬ [CO] [ O 2 ] [ H 2 O ] dt ⎩ RT g ⎭

(69)

Figure 2.5: The rate constants of CO oxidations for four different flame types.

Table 2.7 shows a review of overall rate parameters for CO oxidation for different flame types (experiments). The rate constants for CO oxidation as a function of temperature for 4 different flame types (stirred, jet stirred, turbulent flame and flat flame) are shown in Fig.2.5. There are large differences between them at T = 1000 K and 3000 K. Stirred reactor and flat flame have the same rate constants of CO oxidation at T = 1000 K, while the rate constants of CO oxidation in turbulent flame and jet stirred reactors have the same values. But at T = 3000 K, the rate constants in turbulent flame and flat flame are equal, and stirred and jet stirred reactors have almost the same values.

23

E

Temp.

Pressure

C

[K]

[atm]

[-]

[kcal/mole ]

a

b

c

[-]

[-]

[-]

Reactor

Ref.

11

16.00

1

0.5

0.30

Stirred

[20,36]

12

28.30

1

0.5

0.25

Jet-stirred

[20,49]

Laminar-flow

[20,46]

1280-1535

0.25 - 1

1.20x10

1063-1593

1

1.80x10 12

2.5

*

970-1370

1

1.04x10 /T

32.00

1

0.5

0.25

1030-1230

1

3.90x10

14

40.00

1

0.5

0.25

13

22.66

1

0.5

0.25

14

30.00

1

0.5

0.50

Flat-flame

[37]

13

25.00

1

0.5

0.50

Jet-mixed

[37]

1750-2000

4.00x10

840-2360

1

1.30x10

1400-1800

0.3-0.8

1.80x10

1910-2400

1

1123-1298

8

2.70x10

27.00

10

1.50x10

1

23.6

1

Turbulent

[20,34]

flow

[11,28]

Laminar

0.2,

Bunsen

0.00

**

1

[37,81]

flame

0.5

0.25

Table 2.7: Summary of overall rate parameters for oxidation of carbon monoxide. a

(d[CO]/dt = - C exp{-E/RT} [CO]

b

[H2O]

c

[O2]

-3

[moles cm -3

Concentrations are expressed as mole fraction, not as mole cm .

-1

sec ]).

*

:

**

: 0.2 for [O2] >

5%, 1 for [O2] < 5%

2.3.3 Multistep reaction models Various multistep (chain) reactions for methane, carbon monoxide and hydrogen have been investigated by many authors. Paczko et al. (1986), Seshadri and Peters (1990) have developed a reduced kinetic model with 4 reactions (Eqs.(70)-(73)) for CH4 oxidation from the elementary kinetic model with more than 70 reactions.

CH 4 + 2H + H 2 O → CO + 4 H 2 CO + H 2 O → CO 2 + H 2 2H + M → H 2 + M O 2 + 3 H 2 → 2H + 2 H 2 O

(70) (71) (72) (73)

A multistep global modelling concept for paraffin oxidation has been investigated by Hautman et al. (1981) (Eqs.(74)-(77)). The model adequately reproduced the major species profiles in the flow reactor and characteristic times for shock-tube oxidation. However, they can gives problems with computation due to their negative exponents (Eqs.(78)-(81)).

C n H 2n+2

n C2 H 4 + H 2 2 → 2CO + 2 H 2



C 2 H 4 + O2

24

(74) (75)

1 CO + O 2 → CO 2 2 1 H 2 + O2 → H 2 O 2

(76) (77)

and the rate constants are

⎧ 49600 ⎫ d[ C n H 2n+2 ] 0.50 1.07 0.4 = - 2.09x 1017 exp⎨⎬ [ C n H 2n+2 ] [ O 2 ] [ C 2 H 4 ] dt ⎩ RTg ⎭

(78)

⎧ 50000 ⎫ d[ C 2 H 4 ] 0.90 1.18 -0.37 = - 5.01x 1014 exp⎨⎬ [ C 2 H 4 ] [ O2 ] [ C 2 H 6 ] dt ⎩ RTg ⎭

(79)

⎫ ⎧ 40000 ⎫ d[CO] ⎧⎪ 1.0 0.25 0.5 ⎪ = ⎨- 3.98x 1014 exp⎨⎬ [CO ] [ O 2 ] [ H 2 O ] ⎬ xS dt ⎪⎩ ⎪⎭ ⎩ RTg ⎭

(80)

⎧ 41000 ⎫ d[ H 2 ] 0.85 1.42 -0.56 = - 3.31x 1013 exp⎨⎬ [ H 2 ] [ O2 ] [ C 2 H 4 ] dt R Tg ⎭ ⎩

(81)

where S = max(1.0, 7.93 exp{-2.48φ}), φ is the initial equivalence ratio and S cannot take values greater than 1. Several other multistep reaction models have been studied, e.g. four-step global reaction model for combustion of hydrocarbon (Jones and Lindstedt (1988)); multistep reactions models for CH4 oxidation, e.g. 7 reactions (Enikolopyan (1959)), 18 reactions (Rossberg (1956)), 23 reactions (Boni and Penner (1977)). Multistep reaction models for H2 oxidation have also been studied by others, e.g. 4 reactions (Baulch et aol. (1972), Brown et al. (1974)). Westbrook and Dryer (1981b) described the oxidation mechanism for carbon monoxide by 4 reactions (Eqs.(82)-(85)). The reaction with HO2 is usually negligible in comparison to the reaction with OH (Atri et al. (1977)). Several authors have used different reactions for calculating CO oxidation in methane flames, e.g. Smoot et al. (1976b), Tsatsaronis (1978) used two reactions (Eqs.(82) and (84)), and Jachimowski (1974) used two other reactions (Eqs.(83) and (84))

CO + OH → CO 2 + H CO + O 2 → CO 2 + O CO + O + M → CO 2 + M CO + HO 2 → CO 2 + OH Some rate constants for reaction CO+OH=CO2+H are shown in Table 2.8

25

(82) (83) (84) (85)

Ref.

kCO+OH

11

5.6x10

exp{-1080/RT}

7

1.3

1.51x10 T 11

3.1x10 4.8x10

Baulch and Drysdale (1974) Seery and Bowman (1970)

exp{-700/RT}

Browne et al. (1969)

exp{-5700/RT}

12

2.32x10

exp{758/RT}

exp{-596/RT}

11

3.72x10

-12

Singh and Sawyer (1971)

Kondratiev. (1958)

exp{-5700/RT}

Vandooren et al. (1975)

Table 2.8: The rate constants for reaction CO+OH=CO2+H.

2.3.4 The kinetic reaction models used in the calculations A global model with 3 reactions, 4 reduced models with 10 to 17 reactions, and an elementary model with 123 reactions were studied. The elementary model describes the oxidations of C2H6, CH4, CO, H2 and HCN, while other models describe the oxidations of CO, H2 and CH4. The rate constants for the l reactions are expressed as

⎛ El ⎞ ⎟ k l = Al T βg l exp ⎜⎜ ⎟ ⎝ RTg ⎠

(86)

The units of reaction rates, Al and El, are in moles, cubic centimetres, seconds, Kelvins and calories/mole.

a) Global models (3 reactions): Three reactions (Eqs.(87)-(89)) describing CH4, CO and H2 oxidations were chosen for the calculation of the volatiles evolving from coal

CH 4 + 2 O 2 → CO 2 + 2 H 2 O 1 CO + O 2 → CO 2 2 1 H 2 + O2 → H 2 O 2

(87) (88) (89)

The rate constant for CH4 oxidation is taken from Dryer and Glassman (1973) (Eq.57), while the rate constant for CO oxidation is taken from Howard et al. (1973) (Eq.69). It is difficult to find a global reaction for H2 oxidation. The rate constant used is taken from Hautman et al. (1981), where [C2H4]

- 0.56

is assumed to be unity

⎧ 41000 ⎫ d[ H 2 ] 0.85 1.42 = - 3.31x 1013 exp⎨⎬ [ H 2 ] [ O2 ] dt ⎩ RTg ⎭

26

(90)

b) Reduced model 1 (10 reactions): Ten reactions include a multistep model with 8 reactions describing H2 oxidation, and two reactions for CO and CH4 oxidations as in the global model.

1. 2. 3. 4. 5.

H2+OH=H2O+H H2+O=H+OH H+O2=O+OH OH+OH=O+H2O H+O+M=OH+M H2O/5.0/ 6. O+O+M=O2+M 7. H+H+M=H2+M 8. H+OH+M=H2O+M H2O/20.0/

βl El Al 1.17E9 1.30 3626. 1.80E10 1.00 8826. 5.13E16 -0.82 16507. 6.00E+8 1.30 0. 6.20E16 -0.60 0. 7.50E23 0.00 2.20E18 -1.00 7.50E23 -2.60

0. 0. 0.

(Dixon-Lewis (1968)) (Miller et al. (1984)) (Miller et al. (1982)) (Miller et al. (1984)) (Dixon-Lewis (1968)) (Kee et al. (1985)) (Edelman and Fortune (1969) (Glarborg et al. (1992))

c) Reduced model 2 (10 reactions): The global (irreversible) reaction for CO oxidation in the reduced model 1 is replaced by a reversible reaction, i.e. CO reacts with OH (reaction 1). CH4 oxidation is still calculated by a global reaction as in the global model. Reactions 1., 2., 3. and 4. have been examined in Howard et al. (1973).

1. 2. 3. 4. 5. 6.

CO+OH=CO2+H H2+OH=H2O+H H2+O=H+OH H+O2=O+OH OH+OH=O+H2O H+O+M=OH+M H2O/5.0/ 7. O+O+M=O2+M 8. H+H+M=H2+M 9. H+OH+M=H2O+M H2O/20.0/

Al 1.51E7 1.17E9 1.80E10 5.13E16 6.00E+8 6.20E16

βl 1.30 1.30 1.00 -0.82 1.30 -0.60

El -758. 3626. 8826. 16507. 0. 0.

7.50E23 0.00 2.00E18 -1.00 7.50E23 -2.60

0. 0. 0.

(Baulch et al. (1974)) (Dixon-Lewis (1968)) (Miller et al. (1984)) (Miller et al. (1982)) (Miller et al. (1984)) (Dixon-Lewis (1968)) (Kee et al. (1985)) (Edelman and Fortune (1969) (Glarborg et al. (1992))

d) Reduced model 3 (15 reactions): CH4 oxidation in reduced model 2 is calculated by 6 reversible reactions (reactions 10-15).

1. 2. 3. 4. 5. 6.

CO+OH=CO2+H H2+OH=H2O+H H2+O=H+OH H+O2=O+OH OH+OH=O+H2O H+O+M=OH+M H2O/5.0/ 7. O+O+M=O2+M 8. H+H+M=H2+M 9. H+OH+M=H2O+M

Al 1.51E7 1.17E9 1.80E10 5.13E16 6.00E+8 6.20E16

βl 1.30 1.30 1.00 -0.82 1.30 -0.60

El -758. 3626. 8826. 16507. 0. 0.

7.50E23 2.20E18 7.50E23

0.00 -1.00 -2.60

0. 0. 0.

27

(Baulch et al. (1974)) (Dixon-Lewis (1968)) (Miller et al. (1984)) (Miller et al. (1982)) (Miller et al. (1984)) (Dixon-Lewis (1968)) (Kee et al. (1985)) (Edelman and Fortune (1969) (Glarborg et al. (1992))

10. 11. 12. 13. 14. 15.

H2O/20.0/ CH4+H=CH3+H2 CH4+O=CH3+OH CH4+OH=CH3+H2O CH3+H+M=CH4+M CH3+O=CH2O+H CH3+OH=CH2O+H2

2.20E4 1.60E6 1.60E6 8.00E26 6.80E13 1.00E12

3.00 2.36 2.10 -3.00 0.00 0.00

8750. 7400. 2460. 0. 0. 0.

(Kee (Kee (Kee (Kee (Kee (Kee

et et et et et et

al. al. al. al. al. al.

(1985)) (1985)) (1985)) (1985)) (1985)) (1985))

e) Reduced model 4 (17 reactions): Reduced model 4 is reduced model 3 with two extra rections for CO oxidation.

1. 2. 3. 4. 5. 6. 7. 8.

CO+O+M=CO2+M CO+OH=CO2+H CO+O2=CO2+O H2+OH=H2O+H H2+O=H+OH H+O2=O+OH OH+OH=O+H2O H+O+M=OH+M H2O/5.0/ 9. O+O+M=O2+M 10. H+H+M=H2+M 11. H+OH+M=H2O+M H2O/20.0/ 12. CH4+H=CH3+H2 13. CH4+O=CH3+OH 14. CH4+OH=CH3+H2O 15. CH3+H+M=CH4+M 16. CH3+O=CH2O+H 17. CH3+OH=CH2O+H2

Al 3.20E13 1.51E7 1.60E13 1.17E9 1.80E10 5.13E16 6.00E+8 6.20E16

βl 0.00 1.30 0.00 1.30 1.00 -0.82 1.30 -0.60

El -4200. -758. 41000. 3626. 8826. 16507. 0. 0.

7.50E23 2.20E18 7.50E23

0.00 -1.00 -2.60

0. 0. 0.

2.20E4 1.60E6 1.60E6 8.00E26 6.80E13 1.00E12

3.00 2.36 2.10 -3.00 0.00 0.00

8750. 7400. 2460. 0. 0. 0.

(Kee et al. (1985)) (Baulch et al. (1974)) (Kee et al. (1985)) (Dixon-Lewis (1968)) (Miller et al. (1984)) (Miller et al. (1982)) (Miller et al. (1984)) (Dixon-Lewis (1968)) (Kee et al. (1985)) (Edelman and Fortune (1969) (Glarborg et al. (1992)) (Kee (Kee (Kee (Kee (Kee (Kee

et et et et et et

al. al. al. al. al. al.

(1985)) (1985)) (1985)) (1985)) (1985)) (1985))

f) Elementary model (123 reactions): C2H6, CH4, CO, H2 and HCN oxidations are calculated by 123 reversible reactions. ***************************************** *** REACTION MECHANISM FOR C2H6 *** ***************************************** βl El Al 1. C2H6=CH3+CH3 2.24E19 -1.00 88310. (Westbrook 2. C2H6+CH3=C2H5+CH4 0.55E0 4.00 8280. (Westbrook 3. C2H6+H=C2H5+H2 5.37E2 3.50 5200. (Westbrook 4. C2H6+OH=C2H5+H2O 8.71E9 1.05 1810. (Westbrook 5. C2H6+O=C2H5+OH 2.51E13 0.00 6360. (Westbrook 6. C2H5+M=C2H4+H+M 2.00E15 0.00 30000. (Westbrook 7. C2H5+O2=C2H4+HO2 1.00E12 0.00 5000. (Westbrook 8. C2H4+C2H4=C2H5+C2H3 5.01E14 0.00 64700. (Westbrook 9. C2H4+M=C2H2+H2+M 9.33E16 0.00 77200. (Westbrook 10. C2H4+M=C2H3+H+M 6.31E18 0.00 108720. (Westbrook 11. C2H4+O=CH3+HCO 3.31E12 0.00 1130. (Westbrook 12. C2H4+O=CH2O+CH2 2.51E13 0.00 5000. (Westbrook

28

and and and and and and and and and and and and

Pitz Pitz Pitz Pitz Pitz Pitz Pitz Pitz Pitz Pitz Pitz Pitz

(1984)) (1984)) (1984)) (1984)) (1984)) (1984)) (1984)) (1984)) (1984)) (1984)) (1984)) (1984))

13. C2H4+H=C2H3+H2 14. C2H4+OH=C2H3+H2O 15. C2H4+OH=CH3+CH2O 16. C2H3+M=C2H2+H+M 17. C2H3+O2=C2H2+HO2 18. C2H2+M=C2H+H+M 19. C2H2+O2=HCO+HCO 20. C2H2+H=C2H+H2 21. C2H2+OH=C2H+H2O 22. C2H2+OH=CH2CO+H 23. C2H2+O=C2H+OH 24. C2H2+O=CH2+CO 25. C2H+O2=HCO+CO 26. C2H+O=CO+CH (1984))

1.51E7 2.00 6000. 4.79E12 0.00 1230. 2.00E12 0.00 960. 7.94E14 0.00 31500. 1.00E12 0.00 10000. 1.00E14 0.00 114000. 3.98E12 0.00 28000. 2.00E14 0.00 19000. 6.03E12 0.00 7000. 3.24E11 0.00 200. 3.23E15 -0.60 17000. 6.76E13 0.00 4000. 1.00E13 0.00 7000. 5.01E13 0.00 0.

(Westbrook and Pitz (Westbrook and Pitz (Westbrook and Pitz (Westbrook and Pitz (Westbrook and Pitz (Westbrook and Pitz (Westbrook and Pitz (Westbrook and Pitz (Westbrook and Pitz (Westbrook and Pitz (Westbrook and Pitz (Westbrook and Pitz (Westbrook and Pitz (Westbrook

(1984)) (1984)) (1984)) (1984)) (1984)) (1984)) (1984)) (1984)) (1984)) (1984)) (1984)) (1984)) (1984)) and Pitz

*********************************************** *** REACTION MECHANISME FOR CH4, H2, CO *** *********************************************** 27. CH4+O2=CH3+HO2 7.90E13 0.00 56000. (Kee et al. (1985)) 28. CH4+H=CH3+H2 2.20E4 3.00 8750. (Kee et al. (1985)) 29. CH4+O=CH3+OH 1.60E6 2.36 7400. (Kee et al. (1985)) 30. CH4+OH=CH3+H2O 1.60E6 2.10 2460. (Kee et al. (1985)) 31. CH3+H+M=CH4+M 8.00E26 -3.00 0. (Kee et al. (1985)) 32. CH3+O=CH2O+H 6.80E13 0.00 0. (Kee et al. (1985)) 33. CH3+OH=CH2O+H2 1.00E12 0.00 0. (Kee et al. (1985)) 34. CH3+OH=CH2+H2O 1.50E13 0.00 5000. (Kee et al. (1985)) 35. CH3+H=CH2+H2 9.00E13 0.00 15100. (Kee et al. (1985)) 36. CH2+H=CH+H2 1.40E19 -2.00 0. (Kee et al. (1985)) 37. CH2+OH=CH2O+H 2.50E13 0.00 0. (Kee et al. (1985)) 38. CH2+OH=CH+H2O 4.50E13 0.00 3000. (Kee et al. (1985)) 39. CH+O=CO+H 5.70E13 0.00 0. (Kee et al. (1985)) 40. CH+O2=HCO+O 3.30E13 0.00 0. (Kee et al. (1985)) 41. CH+OH=HCO+H 3.00E13 0.00 0. (Kee et al. (1985)) 42. CH+CO2=HCO+CO 3.40E12 0.00 690. (Kee et al. (1985)) 43. CH2+CO2=CH2O+CO 1.10E11 0.00 1000. (Kee et al. (1985)) 44. CH2+O=CO+H+H 3.00E13 0.00 0. (Kee et al. (1985)) 45. CH2+O=CO+H2 7.80E13 0.00 0. (Kee et al. (1985)) 46. CH2+O2=CO2+H+H 1.60E12 0.00 1000. (Kee et al. (1985)) 47. CH2+O2=CH2O+O 5.00E13 0.00 9000. (Kee et al. (1985)) 48. CH2+O2=CO2+H2 6.90E11 0.00 500. (Kee et al. (1985)) 49. CH2+O2=CO+H2O 1.90E10 0.00 -1000. (Kee et al. (1985)) 50. CH2+O2=CO+OH+H 8.60E10 0.00 -500. (Kee et al. (1985)) 51. CH2+O2=HCO+OH 4.30E10 0.00 -500. (Kee et al. (1985)) 52. CH2O+OH=HCO+H2O 3.43E9 1.18 -447. (Kee et al. (1985)) 53. CH2O+H=HCO+H2 2.19E8 1.77 3000. (Kee et al. (1985)) 54. CH2O+M=HCO+H+M 3.31E16 0.00 81000. (Kee et al. (1985)) 55. CH2O+O=HCO+OH 1.81E13 0.00 3082. (Kee et al. (1985)) 56. HCO+OH=CO+H2O 5.00E12 0.00 0. (Kee et al. (1985)) 57. HCO+M=H+CO+M 1.60E14 0.00 14700. (Kee et al. (1985)) 58. HCO+H=CO+H2 4.00E13 0.00 0. (Kee et al. (1985)) 59. HCO+O=CO2+H 1.00E13 0.00 0. (Kee et al. (1985))

29

60. 61. 62. 63. 64. 65. 66. 67. 68. 69.

HCO+O2=HO2+CO 3.30E13 -0.40 0. (Kee et al. (1985)) CO+O+M=CO2+M 3.20E13 0.00 -4200. (Kee et al. (1985)) CO+OH=CO2+H 1.51E7 1.30 -758. (Baulch and Drysdale (1974)) CO+O2=CO2+O 1.60E13 0.00 41000. (Kee et al. (1985)) CO+HO2=CO2+OH 5.80E13 0.00 22934. (Kee et al. (1985)) H2+O2=OH+OH 1.70E13 0.00 47780. (Kee et al. (1985)) H2+OH=H2O+H 1.17E9 1.30 3626. (Dixon-Lewis (1968)) H2+O=H+OH 1.80E10 1.00 8826. (Miller et al. (1984)) H+O2=O+OH 5.13E16 -0.82 16507. (Kee et al. (1985)) H+O2+M=HO2+M 2.10E18 -1.00 0. (Slack (1977)) H2O/21.0/ H2/3.3/ O2/0.0/ N2/0.0/ 70. H2+M=H+H+M 2.23E12 0.50 92600. (Kee et al. (1985)) H2O/6.0/ H/2.0/ H2/3.0/ 71. H+H+H2=H2+H2 9.20E16 -0.60 0. (Kee et al. (1985)) 72. H+H+H2O=H2+H2O 6.00E19 -1.25 0. (Kee et al. (1985)) 73. H+H+CO2=H2+CO2 5.49E20 -2.00 0. (Kee et al. (1985)) 74. H+OH+M=H2O+M 7.50E23 -2.60 0. (Glarborg et al. (1992)) H2O/20.0/ 75. H+O+M=OH+M 6.20E16 -0.60 0. (Dixon-Lewis (1968)) H2O/5.0/ 76. H+HO2=H2+O2 2.50E13 0.00 700. (Miller et al. (1984)) 77. HO2+HO2=H2O2+O2 2.00E12 0.00 0. (Miller et al. (1984)) 78. HO2+H=OH+OH 2.50E14 0.00 1900. (Miller et al. (1984)) 79. H2O2+M=OH+OH+M 1.30E17 0.00 45500. (Baulch et al. (1972)) 80. H2O2+H=HO2+H2 1.60E12 0.00 3750. (Baulch et al. (1972)) 81. H2O2+OH=H2O+HO2 1.00E13 0.00 1800. (Baulch et al. (1972)) 82. OH+HO2=H2O+O2 5.00E13 0.00 0. (Miller et al. (1984)) 83. O+HO2=O2+OH 4.80E13 0.00 1000. (Miller et al. (1984)) 84. OH+OH=O+H2O 6.00E+8 1.30 0. (Miller et al. (1984)) *************************************** *** REACTION MECHANISM FOR HCN *** *************************************** 85. HCN+OH=CN+H2O 4.40E12 0.00 9000. 86. HCN+O=NCO+H 1.21E4 2.64 4980. 87. HCN+O=NH+CO 5.17E3 2.64 4980. 88. HCN+O=CN+OH 2.70E9 1.58 26600. 89. CN+H2=HCN+H 5.45E11 0.70 4885. 90. CN+O=CO+N 1.80E13 0.00 0. 91. CN+O2=NCO+O 5.60E12 0.00 0. 92. CN+OH=NCO+H 5.00E13 0.00 0. 93. NCO+H=NH+CO 5.00E13 0.00 0. 94. NCO+O=NO+CO 3.00E13 0.00 0. 95. NCO+N=N2+CO 2.00E13 0.00 0. 96. NCO+OH=NO+CO+H 1.00E13 0.00 0. 97. NCO+M=N+CO+M 3.10E16 -0.50 48000. 98. NCO+NO=N2O+CO 1.90E13 0.00 0. 99. NCO+H2=HNCO+H 8.58E12 0.00 9000. 100. HNCO+H=NH2+CO 2.00E13 0.00 3000. 101. NH+O2=HNO+O 1.00E7 0.00 12000. 102. NH+O2=NO+OH 1.40E11 0.00 2000. 103. NH+NO=N2O+H 4.33E14 -0.50 0.

30

(Miller (Miller (Miller (Miller (Miller (Miller (Miller (Miller (Miller (Miller (Miller (Miller (Miller (Miller (Miller (Miller (Miller

et al. (1984)) et al. (1984)) et al. (1984)) et al. (1984)) et al. (1984)) et al. (1984)) (Miller et al. (1984)) et al. (1984)) et al. (1984)) et al. (1984)) (Miller et al. (1984)) et al. (1984)) et al. (1984)) et al. (1984)) et al. (1984)) et al. (1984)) et al. (1983)) et al. (1983)) et al. (1983))

104. 105. 106. 107. 108. 109. 110. 111. 112. 113. 114. 115. 116. 117. 118. 119. 120. 121. 122. 123.

N2O+H=N2+OH NH+OH=HNO+H NH+OH=N+H2O NH+N=N2+H NH+H=N+H2 NH2+O=HNO+H NH2+O=NH+OH NH2+OH=NH+H2O NH2+H=NH+H2 NH2+NO=NNH+OH NH2+NO=N2+H2O NNH+M=N2+H+M NNH+NO=N2+HNO NNH+H=N2+H2 HNO+M=H+NO+M HNO+OH=NO+H2O HNO+H=H2+NO N+NO=N2+O N+O2=NO+O N+OH=NO+H

7.60E13 0.00 15200. (Miller 2.00E13 0.00 0. (Miller 5.00E11 0.50 2000. (Miller 3.00E13 0.00 0. (Miller 3.00E13 0.00 0. (Miller 6.63E14 -0.50 0. (Miller 6.75E12 0.00 0. (Miller 4.50E12 0.00 2200. (Miller 6.92E13 0.00 3650. (Miller 8.82E15 -1.25 0. (Miller 3.78E15 -1.25 0. (Miller 2.00E14 0.00 20000. (Miller 5.00E13 0.00 0. (Miller 3.70E13 0.00 3000. (Miller 1.50E16 0.00 48680. (Miller 3.60E13 0.00 0. (Miller 5.00E12 0.00 0. (Miller 3.27E12 0.30 0. (Miller 6.40E9 1.00 6280. (Miller 3.80E13 0.00 0. (Miller

31

et et et et et et et et et et et et et et et et et et et et

al. al. al. al. al. al. al. al. al. al. al. al. al. al. al. al. al. al. al. al.

(1983)) (1983)) (1983)) (1984)) (1983)) (1983)) (1983)) (1983)) (1983)) (1983)) (1983)) (1983)) (1983)) (1983)) (1983)) (1983)) (1983)) (1984)) (1984)) (1984))

2.4 Combustion of coal char The overall reaction process involves several steps in sequence: transport of oxygen or other reactant gas to the surface of the particle, reaction with the surface, and transport of products away. Oxygen is transported to the surface and products are transported away by molecular diffusion, owing to concentration gradients between the surface and the free gas stream. The most important reactions on the surface of particles may be (Field et al. (1967))

C f + O 2 → CO 2 1 C f + O 2 → CO 2 C f + O → CO C f + CO 2 → 2CO C f + H 2 O → CO + H 2

(91) (92) (93) (94) (95)

Cf is the fixed carbon of coal char. Depending on the nature of products formed at the particle surface, several char oxidation models are proposed: single-film model, double- film model and continuous-film model. - Single-film model: (Caram and Amundson (1977), Laurendeau (1978), Mitchell and Madsen (1986)) The particle is consumed via reactions with oxygen and no reaction occurs in the boundary layer. CO and CO2 are the heterogeneous products and CO reacts with oxygen after it has been transported to the free stream. If the particle temperature is above 1273 K, the main product of the reaction is CO (Arthur (1951)). The single-film model is applicable for particle diameters less than 100 µm (Ayling and Smith (1972)). - Double-film model: (Caram and Amundson (1977), Laurendeau (1978), Mitchell and Madsen (1986))

Oxygen from the gas phase diffuses to the surface of the particle and

reacts with the fixed carbon of the coal to form carbon monoxide (CO) and carbon dioxide (CO2) (Eqs.(1) and (2)). One-half of the CO2 formed in the reaction zone surrounding the particles diffuses back to the particle surface and reacts with the fixed carbon to form carbon monoxide (Eq.(4)) and the other one-half diffuses out into the main air stream. The CO formed diffuses outwards into the air stream where it reacts with the incoming oxygen to form carbon dioxide in the gas phase. The double film is adequate for particle diameters greater than 1 mm (Wicke and Wurzbacher (1962)). - Continuous-film model: (Caram and Amundson (1977), Laurendeau (1978), Mitchell and Madsen (1986) This model is important for particle sizes between 100 and 1000 µm. The oxygen of air comes in contact with the carbon of the particle, and both CO and CO2 are produced. However, if the temperature is greater than about 923 K (the ignition temperature of CO), the CO burns anywhere within the boundary layer depending on particle size. If the particle temperature is above 1373 K, CO2 is reduced to CO on the particle surface.

32

In this work, the burning of coal char is described by the single-film model, i.e. CO and CO2 are the heterogeneous products and CO reacts with oxygen after it has been transported to the free stream. Two models are used for calculations. The first one, called char oxidation model 1, is described in Mitchell (1988) and the second, or simple one, called char oxidation model 2, is described in Field et al. (1967). It is called a simple model because the apparent reaction order equals unity. We must distinguish between two different meanings for the word "model" in this subsection: "model" in the

single-film

model, double-film model and continous-film model indicates the products and how they are formed on the surface of particle; "model" in e.g. model of Field et al. (1967) and Mitchell (1988) indicates the speed at which the particles can burn.

2.4.1 Char oxidation model 1 (Mitchell's model): (Mitchell (1988)) The overall particle burning rate per unit external surface area is expressed as

⎧ γ Ps ⎫ ⎪⎪ 1 - P ⎪⎪ K d q= ln ⎨ ⎬ γ ⎪1 - γ P g ⎪ ⎪⎩ P ⎪⎭

(96)

P is the total pressure, Ps is the oxygen partial pressure at the particle surface, Pg is the oxygen partial pressure in the gas phase and Kd is the mass transfer (diffusive) coefficient, given by

Kd =

P M c Dox (Sh) d p R′ T m ν o

(97)

γ = (ψ-1)/(ψ+1) is the change in volume during reaction referred to unit volume of oxygen, ψ = 1/(1+CO/CO2) is the fraction of carbon converted to CO2 at the particle surface, v0 = 0.5(1+ψ) is the stoichiometric oxygen coefficient for reaction at the particle surface, Sh is 3

the Sherwood number, R'= 82.06 atm cm

-1

mole

-1

K

is the gas constant, Mc is the

molecular weight of carbon, and Dox is the diffusion coefficient of oxygen. It is defined as (Field et al. (1967)) 1.75

⎛Tm⎞ Dox = Dox,0 ⎜⎜ ⎟⎟ ⎝ T0 ⎠ 2

Dox,0 = 0.207cm /s for O2-N2 at 300 K. The Sherwood number, Sh, is calculated as

33

(98)

Sh =

λ dp D ki

1

1

= 2.0 + 0.65 (Re )2 (Sc )3 (99)

Sc = v/Dki is the Schmidt number; Sh = 2 for velocities of gas and particles are equal. The apparent density and diameter of a burning char particle are defined a

ρp ⎛ mp ⎞ ⎟ = ⎜⎜ ρ p,0 ⎝ m p,0 ⎟⎠

(100)

and b

⎛ mp ⎞ dp ⎟ = ⎜⎜ ⎟ d p,0 ⎝ m p,0 ⎠

(101)

The subscript 0 denotes the initial values. According to Mitchell (1988), for spherical particles, a + 3b = 1.0. For constant density burning, a = 0 and for constant diameter burning a = 1.0. For values of a between 0 and 1, both the size and density of a particle decrease with burnoff. Further details of a and b can be found in Mitchell (1988), Smith (1971a,b). In terms of an apparent chemical reaction rate coefficient, Ks, and an apparent reaction order with repect to the oxygen partial pressure, n, the overall particle burning rate is expressed as

q = K s P ns

(102)

Ks is the apparent chemical reaction rate coefficient, written as

⎛ Ea ⎞ ⎟ K s = Aa exp ⎜⎜ R T p ⎟⎠ ⎝

(103)

Inserting Ps in Eq.(102) into Eq.(96), yields 1

γ ⎧ q ⎫n

⎛ γ ⎞ Pg ) exp⎜⎜ q ⎟⎟ - 1 = 0 ⎨ ⎬ + (1 - γ P P ⎩Ks⎭ ⎝ Kd ⎠

(104)

or

F(q) =

γ ⎧q⎫

1 n

⎛ γ ⎞ Pg ) exp⎜⎜ q ⎟⎟ - 1 = 0 ⎨ ⎬ + (1 - γ P P ⎩Ks⎭ ⎝ Kd ⎠

Differentiating Eq.(104), yields

34

(105)

1 ∂F(q) 1 P ⎧ ⎛ γq ⎞ Pg ⎫ -1 n + = ⎟⎟ q ⎨1 - γ ⎬ exp⎜⎜ 1/n ∂q n Ks P⎭ Kd ⎩ ⎝ Kd ⎠

(106)

Eq.(105) and (106) are nonlinear algebraic equations which can be solved by the NewtonRaphson method. The Newton-Raphson method is implemented by executing the following sequence of steps 0

1. Guest q

0

n

1

1

n+1

in

1

n

2

2

n+1

in

2. Setting q = q into Eqs.(105) and (106) then calculating q (q ≈ q Eq.(107))

3. Setting q = q into Eqs. (105) and (106) then calculating q (q ≈ q n+1

Eq.(107)), etc. until q

n

- q < eps (e.g. eps = 1.0E-10). -1

q

n+1

⎧ ∂F(q) ⎫ n =q -⎨ ⎬ • F( q ) ⎩ ∂q ⎭n n

(107)

The combustion parameters for Mitchell's char oxidation model are shown in Table 2.9. The combustion parameters of Mitchell's model are different for different coal types.

Coal type

Pre-

Activation

exponential

energy

order of

Particle

O2-

diameter

environment

Temperature

Ref.

reaction

factor -2

Apparent

-1

s

[cal/mol]

[-]

[µm]

[%]

6907.0

34133.0

0.17

13.0

0.1-0.5

1700-2200

1.72

16790.0

0.25

75-125

6 and 12

1550-1700

[58]

22.4

22500.0

0.50

106-125

6 and 12

1535-1700

[40]

4564.0

33967.0

0.17

3.3

0.1-0.5

1700-2200

[50]

664.0

29194.0

0.17

14.9

0.1-0.5

1700-2200

[50]

29.0

24000.0

0.5

75-125

6 and 12

1535-1700

[40]

7.21

19310.0

0.5

106-125

6 and 12

1550-1700

[58]

9.0

24100.0

0.1

106-125

6 and 12

1535-1700

[40]

-

-

-

-

-

-

-

Ulan

49.4

29000.0

0.2

106-125

6 and 12

1535-1700

[40]

Blair Atholl

11.1

24700.0

0.1

106-125

6 and 12

1535-1700

[40]

[g cm

[K]

-n

atm ] Illinois #6

Pittsburgh #8

Cerrejon Middleburgh

Table 2.9: Summary of char combustion parameters. 2.4.2 Char oxidation model 2 (Field et al. (1967))

35

[50]

The rate of reaction of carbon with oxygen can be expressed as

q = K s Ps

(108) 2

where q is the rate of consumption of carbon (g/(cm s)), Ks is the surface reaction rate 2

coefficient (g/(cm s atm)), and ps is the oxygen partial pressure at the surface (atm). The surface reaction rate coefficient is independent of the oxygen concentration, and depends on the surface temperature as well as on the fuel type. Kd is defined as

K s = Aa exp(-

Ea ) RT p

(109) 2

where Tp is the particle temperature at the surface, Aa = 8710 g/cm , Ea = 35700 cal/mole are the empirical values. Ks is discussed in Field et al. (1967). In the boundary layer, assumed that the oxygen is conserved, i.e. oxygen is transport through a imaginary surface at any radius and is equal to the rate of oxygen transported to the particle surface.

4 π r 02 G( r 0 ) = 4 π r 2 G(r)

(110)

where r is the radial distance from the centre of the particle, r0 is the radius of the particle, and G(r) is the flux of oxygen at radius r. The flux of oxygen is given by the diffusional transport equation

G(r) = -

Dox M o dP R′ T g dr

(111)

Dox is the diffusion coefficient of oxygen in the gas, Mo is the molecular weight of oxygen, R' 3

is the gas constant (= 82.06 atm cm /(mole K)), T is the gas temperature, and P is partial pressure of oxygen. Substituting for the oxygen flux in Eq.(110) yields 2 r 0 G( r 0 ) =

or 2

r 0 G( r 0 )

Dox M o 2 dP r dr R′ T g

R′ T g 1 Dox M o r

36

2

=

dP dr

(112)

(113)

Assuming that the variation of temperature can be disregarded, integrating Eq.(113) between any radius r and the particle surface r0 , yields

r 0 G( r 0 )

R′ T m ⎛ r 0 ⎞ ⎜ 1 - ⎟ = P(r) - P( r 0 ) Dox M o ⎝ r ⎠

(114)

Far from the particle, r0/r approaches zero and the oxygen concentration tends to be the value in the free stream, i.e.

r 0 G( r 0 )

R′ T m Dox M o

= P g - Ps (115)

Pg and Ps are the oxygen partial pressures in the free stream and at the particle surface, respectively. This equation gives the rate of transport of oxygen to the particle surface in terms of the concentration of oxygen in the free stream and at the surface. The ratio of carbon consumed to oxygen transported to the surface depends on the product which is transported away. In general, the relation may be written

q

=

ϕ G( r 0 )

Mc or

Mo

(116)

ϕ G( r 0 ) q = Mc Mo

(117)

2

where q is the rate of consumption of carbon (g/(cm s)), _ is the mechanism factor which takes the values 1 when CO2 is the sole heterogeneous product and 2 when CO is the sole product. Combining Eqs.(115) and (117) yields

where

ϕ q = M c Dox ( P g - P s ) = K d ( P g - P s ) r 0 R′ T m

(118)

M c ϕ Dox Kd = r 0 R′ T m

(119)

Kd is the diffusional reaction rate coefficient. It depends on the particle diameter, the mechanism factor and the mean temperature, but is completely independent of fuel type. Eq.(97) is similar to Eq.(119) with P = 1 atm, Sh = 2.0 and v0 = 1/_. In all calculations, Kd is taken from Eq.(97).

37

The mean temperature can be calculated as

Tm =

1 (T g +T p ) 2

(120)

If the surface reaction rate coefficient is sufficiently high, the oxygen partial pressure at the surface may be small compared with the value in the free stream. The reaction is then said to be diffusion-controlled. Combining Eqs.(108) and (118), we obtain

q=

pg 1 Ks

+

1 Kd

(121)

This equation is used to calculate the rate of consumption of carbon. Ks and Kd are described in Eqs.(109) and (119), respectively. 2.4.3 Formation of CO and CO2 The model allows for both CO and CO2 formation at the particle surface, but assumes that no reaction occurs in the boundary layer surrounding the particle. According to Mitchell [1988], the ratio CO/CO2 increases with increasing temperature. Formation of CO and CO2 can be expressed as

⎛ ⎞ (moles CO) = ACO exp⎜⎜ - E CO ⎟⎟ (moles CO 2 ) ⎝ R* T p ⎠

(122)

where ACO and ECO are the pre-exponential factor and activation energy for formation of CO and CO2, respectively. The empirical constants of ACO and ECO are shown in Table 2.10. In Jensen and Mitchell (1993), ACO and ECO are different for different coal types.

38

ACO (-) Arthur (1951)

11

12400.0

3

14300.0

11

73200.0

2.50 x 10

8

60000.0

4

30000.0

2.512 x 10

Rossberg (1956)

1.995 x 10

Mitchell (1988) Jensen and Mitchell (1993)

ECO (cal/mole)

2.630 x 10 Illinois #6

Pittsburgh #8

4.0 x 10

Cerrejon

*

0.0

*

0.0

*

0.0



Blair Athol



Ulan



*

Table 2.10: Summary of ratio of CO/CO2 product. : CO is the sole heterogeneous reaction product. The ratio of CO to CO2 as a function of the inverse temperature (1/T) is shown in Fig.2.6, and the fraction of carbon converted to CO2 at the particle surface, ψ, as function of temperature is shown in Figs.2.7 and 2.8. The CO/CO2 and ψ profiles calculated by Arthur's and Rossberg's constants are nearly the same, while the CO/CO2 and ψ profiles using Mitchell's constant are quite different.

Figure 2.6: CO/CO2 as function of 1/T (T: particle temperature).

39

Figure 2.7: Mole fraction of CO2 as function of 1/T (T: particle temperature).

Figure 2.8: Mole fraction of CO2 as function of T (T: particle temperature).

40

3 RESULTS AND DISCUSSIONS Numerous mathematical models for coal reaction process, including devolatilization, char oxidation, gas phase oxidation, and gas-particle interchange have been proposed. However, the choice of single subprocess, e.g. devolatilization, char oxidation, gas phase oxidation, is very difficult. In this work, two devolatilization models, Multiple Parallel Reaction Model (MPRM) and Distributed Activation Energy Model (DAEM), were used for decomposition kinetics. Six sets of multistep reaction models were used for calculation of combustion of volatiles (CO, HCN, C2H6, CH4 and H2). The combustion of coal char was described by single film model with two different "reaction rates" called Mitchell's model (model 1) and model of Field et al. (1967) (model 2). The influence of radiative heat transfer between particles and radiation of hot particles to the cold environment were also examinated. The measured gas temperatures were used to compute reaction rates. The gas temperatures are known (measured) so that they could not influence the kinetic (homogeneous) reactions. The temperture difference between the particles and the gas depends on the rate of heat generation (char oxidation) and on the rate at which the particle dissipates heat by conduction (convection if there is slip velocity between gas and particles) and radiation. The heat generation depends on the amount of consumed fixed carbon (over all particle burning rate) to form carbon monoxide and carbon dioxide. Without the char oxidation model in the calculations, the cold particles are only heated by the hot gas (conduction) and hot particles (radiation) and give simultaneously heat to the cold environment by radiation. With the char oxidation model, the cold particles are heated by hot gas and hot particle. At the exit plan of the burner (burner outlet), they begin to release volatiles and ignite (char burning). The particles now are hotter than the surrounding gas, they heat the gas by conduction and heat the cold particles at both ends of the flame by radiation. Illinois #6 and Pittsburgh #8 coals were chosen for testing due to their many available data. Major species profiles (CO2, O2, CO, CH4 and H2) were compared with the experimental data. All calculated mole fractions of species are converted to "dry basic" data. The unit in all concentration figures is written in mole fraction with the 6 major gases N2, CO2, O2, CO, CH4 and H2. The gas temperture profile (GTP) and the particle temperature profile (PTP) from calculations and experiment were also shown. The "calculated" gas temperatures have the same values as the experamental data, and the particle temperatures were calculated by Eq.(2). The coordinate with negative sign means that the distance before burner outlet. The oxidizer is the dry air with mole fractions as follow: %N2 = 78.09, %O2 = 20.95,, %CO2 = 0.03, and %AR = 0.93. The amount of volatiles in calculation were taken as the volatile matters from the proximate analysis times with the Qfactor, which were taken from Jensen and Mitchell (1993) (1.385 for Illinois #6 coal and 1.50 for Pittsburgh #8 coal). Except for the cases called "New ultimate analysis", all analysis values (both proximate and ultimate values) of Illinois #6 and Pittsburgh #8 coals used in the calculations were taken from column (2) of Table A3 (APP.A1). The used analysis values for case "New ultimate analysis" were taken from column (3) of Table A3 (only for Illinois

41

#6 coal). The calculated resisdence time of coal particles in Illinois #6 coal air flame is shown in Fig.3.1, and the calculated gas (particle) velocity profile is shown in Fig.3.2. The resisdence time of coal particles from the burner outlet (x = 0.0 cm) to the end of the calculated domain (x = 3.0 cm) is about 0.046 sec. After the burner outlet, the velocity of gas (particle) increases very fast due to the hot gas (low gas mass density).

Figure 3.1: The calculated residence time of coal particles.

Figure 3.2: The calculated gas (particle) velocity in the Illinois #6 coal air flame.

Table 3.1 shows the compositions of combustion products for six sets of kinetic reaction models and the total mole fractions of major species. The total mole fractions of minor products vary from 8.3% for using global model to 18.0% for using elementary model (by difference with the total mole fraction of major products in Table 3.1). Only CO2, O2, CO, CH4 and H2 concentration (mole fraction) profiles are shown and compared with the experimental data.

42

Volatile combustion models

Number

Composition of combustion

of

products

Sum of mole fractions of O2, H2, CO2,

reactions

CO, N2, CH4 (major products)

Global Model

3

H2O, O2, H2, CO2, CO, C2H6,

Without char

With char

oxidation

oxidation

model (%)

model (%)

91.69

91.90

85.96

85.94

85.97

86.40

88.06

86.59

88.07

87.07

82.06

----

HCN, Ar, N2, CH4 Reduced

10

H2O, O2, O, OH, H2, H, CO2, CO, C2H6, HCN, Ar, N2, CH4

Model 1 Reduced

10

H2O, O2, O, OH, H2, H, CO2, CO, C2H6, HCN, Ar, N2, CH4

Model 2 Reduced

15

H2O, O2, O, OH, H2, H, CO2, CO, C2H6, HCN, Ar, N2, CH4, CH3,

Model 3

CH2O Reduced

17

H2O, O2, O, OH, H2, H, CO2, CO, C2H6, HCN, Ar, N2, CH4, CH3,

Model 4

CH2O Elementary Model

H2O, O2, O, OH, H2, H, CO2, CO, 123

C2H6, HCN, H2S, Ar, N2, CH4, CH3, CH2O, C2H5, C2H4, C2H3, C2H2, C2H, CH2, CH, CH2CO, HCO, HO2, H2O2, HNCO, HNO,CN,NCO, NNH, NH2, NH, N2O, NO, N

Table 3.1: The compositions of combustion product for six sets of kinetic reaction models.

3.1. ILLINOIS #6 COAL AIR FLAME ……………………………………………………………………………………. ………………………………not available……………………………………. …………………………………………………………………………………….

3.2. PITTSBURGH #8 COAL AIR FLAME ……………………………………………………………………………………. ………………………………not available…………………………………….

43

…………………………………………………………………………………….

44

CONCLUSIONS Two devolatilization models, six reaction sets describing combustion of volatiles, and two char oxidation models were used to calculate species concentrations and particle temperatures for Illinois #6 and Pittsburgh #8 coal air flames. The combustion of coal char is described by the single-film model. The influence of radiative heat transfer between particles and a cold environment were also examined. The mass-mean particle sizes are 15.0 µm and 14.0 µm for Illinois #6 and Pittsburgh #8 coal, respectively. Two devolatilization models called the multiple parallel reaction (MPRM) and distributed activation energy model (DAEM) were used for testing. There were 7 gases in the volatiles: CO2, H2O, CO, C2H6, CH4, H2 and HCN. There was no difference between the two models. For high temperatures, the devolatilization process occurs quickly and finishs about 3 - 5 mm (0.06 sec) from the burner. Combustion (oxidation) of volatiles with and without char oxidation were also tested. There was five active gases, CO, C2H6, CH4, H2 and HCN in volatiles, three of which (CO, CH4 and H2) were most significant for coal combustion due to their high activities. Six different sets of kinetic reaction called global model, reduced model 1, 2, 3 and 4, and elementary model were used for testing. They varied from 3 reactions (global model) to 123 reactions (elementary model). All kinetic models described oxidations for CO, CH4 and H2, except the elementary model which described oxidations for all five active gases (CO, C2H6, CH4, H2 and HCN). Without the char oxidation model, CO is the sole product of the pyrolysis process and it is quickly oxidized. All concentrations calculated without char oxidation model were too low compared with the experimental data. The particle temperatures were heated by hot gas by conduction and always lower than gas temperature. With the char oxidation model, CO comes partly from the pyrolysis, partly from the char oxidation. Except for the elementary kinetic reaction model, all kinetic models were calculated with the char oxidation model. The elementary model with 123 reactions was too comprehensive to calculate with the char oxidation model (too many reactions). The concentrations for CO2, O2 and CO calculated with the reduced model 3 with 15 reactions agreed with the experimental data. This case is called the reference case. The amount of CH4 was too small and therefore difficult to compare. The H2 concentration was poor compared with experimental data. This is because is the H2 may come from char oxidation (Eq.(95)) which has not been considered in the calculation. The particle temperatures obtained from this case agreed well with the experimental data. The largest difference between calculation and experiment was about 50.0 K. The error of particle temperature may come from the error of measurement of gas temperature. However, it is acceptable. The burning of coal char is described by the single-film model. Both CO and CO2 are the

45

heterogeneous products. Two different "reaction rates" called Mitchell's model (Model 1) and the model of Fields et al. (1967) (Model 2) were used. Although the results obtained with the char oxidation model 2 were in good agreement with the experimental data, the reaction rate of model 2 seems to be low and therefore the reaction between oxygen from the gas phase and fixed carbon of coal char was slower compared with the experimental data. Mitchell's model, where the apparent reaction order differed from unity, is complicated for calculation. The reaction between fixed carbon and oxygen from the surroundings was also slow compared with that of the experimental data. The model of Field et al. (1967) is suitable for calculating a coal air flame if the computional time is considered. The composition of volatiles (mass fractions of single pyrolysis species), e.g. YCO2, YH2O, YCO, YHCN, YC2H6, YCH4, YH2 and YS (YH2S), are usually unknown and must be estimated. The values are calculated from the measured composition of volatiles of Solomon et al. (1992) (Table A4, APP.A1) and they are adjusted by multiplying with a Q-factor of 1.385 and 1.50 for Illinois #6 and Pittsburgh #8 coals, respectively. The ultimate analysis gives very sensitive results. The pyrolysis process takes place very quickly, and there is therefore only a slight difference between results based on the assumptions that the char oxidation occur after and simultaneously with the pyrolysis process. But for small particles, the amount of gases released from coal by pyrolysis may be too small to prevent the oxygen from the gas phase being transported to the particle surface and reacting with it. Therefore, the assumption that the char oxidation process and the pyrolysis process occur at the same time is more corrected. A knowledge of particle size is very important for calculation. The ignition of the small particles begins earlier and they are burnt faster than the large particles. The temperatures of large particles may be higher than the temperatures of small particles. The concentration profiles for three different particle sizes were also different because of different particle temperatures. Three different oxygen concentrations in the oxidizer are were for testing. More CO2 is produced with higher oxygen concentrations in the oxidizer. The CO2 comes partly from the pyrolysis, partly from the char oxidation, and partly from the oxidation of CO in the gas phase. The CO concentration was also high but it was quickly oxidized in the gas phase. The CO concentration is high with a low oxygen concentration in the oxidizer. The heat generation by char oxidation depends on the oxygen concentration in the oxidizer because experiments were performed with fuel-rich mixture (equivalence ratio Φ = 2.14 for Illinois #6 coal and Φ = 2.09 for Pittsburgh #8 coal). The greater the oxygen concentration in the oxidizer, the greater the char oxidation and the more energy generated. Both the radiation of particles to a cold environment and the radiation between particles influence the results. However, the influence of these radiations is small due to

small

amounts of particles. This influence may be more important with large amounts of particles.

46

The peak of particle temperature profiles is influenced most because the particle temperatures at this point are highest. The concentration profiles for the cases without radiation of particles to a cold environment and radiation between particles were slightly different compared with the reference case, due to different particle temperatures. The results obtained for Illinois #6 coal were closer to the experimental data than those obtained for Pittsburgh #8 coal, because the amount of high hydrocarbon (e.g. C2H6) and (or) tar for Pittsburgh #8 coal was higher than for Illinois #6 coal. The soot formation in Pittsburgh #8 coal air flame may therefore be higher than that in Illinois #6 coal air flame.

47

REFERENCES [1]

Anthony, D.B., Howard, J.B., Hottel, H.C., and Meissner, H.P. (1974). Rapid Devolatilization of Pulverized Coal. 15th Symp. (Int.) on Combust.. The Combust. Inst., PA, pp. 1303-1317.

[2]

Anthony,

D.B.,

and

Howard,

J.B.

(1976).

Coal

Devolatilization

and

Hydrogasification. AIChE Journal, 22, No. 4., pp. 625-656. [3]

Arthur, J. (1951). Trans. Faraday Soc., 47, pp. 164.

[4]

Atri, G.M., Baldwin, R.R., Jackson, D., and Walker, R.W. (1977). The Reaction of OH Radicals and HO2 Radicals with Carbon Monoxide. Combust. and Flame, 30, pp. 1-12.

[5]

Ayling, A.B., and Smith, I.W. (1972). Measured Temperatures of Burning Pulverized-Fuel Particles, and the Nature of the Primary Reaction Product. Combust. and Flame, 18, pp. 173-184.

[6]

Basevich, V.Ya. (1987). Chemical Kinetics in the Combustion Processes: A Detail Kinetics Mechanism and Its Implementation. Prog. Energy Combust. Sci., 13, pp. 119-248.

[7]

Baulch, D.L., Drysdale, D.D., Horne, D.G., and Lloyd, A.C. (1972). Evaluated Kinetic Data for High Temperature Reactions. Volume 1. Homogeneous Gas Phase Reactiosn of the H2-O2 Systems. London Butterworths.

[8]

Baulch, D.L., and Drysdale, D.D. (1974). An Evaluation of the Rate Data for the Reaction CO + OH ---> CO2 + H. Combust. and Flame, 23, pp. 215-225.

[9]

Baulch, D.L., Coboc, C.J., Cox, R.A., Frank, P., Hayman, G., Just, Th., Kerr, J.A., Murrells, T., Pilling, M.J., Troe, J., Walker, R.W., and Warnatz, J. (1994). Summary Table of Evaluated Kinetic Data for Combustion Modeling: Supplement 1. Combust. and Flame, 98, pp. 59-79.

[10]

Boni, A.A., and Penner, R.C. (1977). Sensitivity Analysis of a Mechanism for Methane Oxidation Kinetics. Combust. Sci. Technol., 15, pp. 99-106.

[11]

Browman, C.T. (1975). Kinetics of Pollutant Formation and Destruction in Combustion. Prog. Energy Combust. Sci., 1, pp. 33-45.

[12]

Brown, N.J., Fristrom, R.M., and Sawyer, R.F. (1974). A Simple Premixed Flame Model Including an Application to H2 + Air Flames. Combust. and Flame, 23, pp. 269-275.

[13]

Browne, W.G., White, D.R., and Smookler, G.L. (1969). A Study of the Chemical Kinetics of Shock Heated. 12th Symp. (Int.) on Combust., The Combust. Inst., PA, pp. 557-567.

[14]

Caram, H.S., and Amundson, N.R. (1977). Diffusion and Reaction in a Stagnant Boundary Layer about a Carbon Particle. Ind. Eng. Chem., Fundam., 16, No. 2, pp. 171-181.

[15]

Carangelo, R.M., Solomon, P.R., and Gerson, D.G. (1987). Application of TGFT-ir to study hydrocarbon structure and kinetics. Fuel, 66, pp. 960-967.

[16]

Carpenter, A.M., and Skorupska, N.M. (1993). Coal combustion-analysis and testing. IEACR/64, IEA Coal Research, London.

[17]

Chapman, S., and Cowling, T.G. (1970). The Mathematical Theory of Non-

48

Uniform Gases. Third Edition, Cambridge University Press, Cambridge. [18]

Curtiss, C.F., and Hirschfelder, J.O. (1949). Transport Properties of Multicomponent Gas Mixtures. J Chem Phys., 17, pp. 550-.

[19]

Dixon-Lewis, G. (1968). Flame Structure and Flame Reaction Kinetics, II. Transport Phenomena in Multicomponent Systems, Proc. Roy. Soc., A 307, pp. 111-135

[20]

Dryer, F.L., and Glassman, I. (1973). High-Temperature Oxidation of CO and CH4. 14th Symp. (Int.) on Combust., The Combust. Inst., PA, pp. 987-1003.

[21]

Dryer, F.L. (1991). The Phenomenology of Modeling Combustion Chemistry. In Fossil Fuel Combustion. A Source Book. Edited by Bartok, W. and Sarofim, A.F.. Joln Wiley & Sons, Inc., pp. 121-213.

[22]

Edelman, R.B., and Forune, O.F. (1969). A Quasi-Global Chemical Kinetic Model for the Finite Rate Combustion of Hydrocarbon Fuels with Application to Turbulent Burning and Mixing in Hypersonic Egines and Nozzles. AIAA Paper, No. 69-86, pp. 1-11.

[23]

Enikolopyan, N.S. (1959). Kinetics and Mechanism of Methane Oxidation. 7th Symp. (Int.) on Combust., The Combust. Inst., PA, pp. 157-164.

[24]

Essenhigh, R.H. (1981). Fundamental of Coal Combustion. In Chemistry of Coal Utilization, Second Supplementary Volume. Edited by Elliott, M.A., John Wiley & sons, Inc., pp. 1153-1312.

[25]

Field, M.A., Gill, D.W., Morgan, B.B., and Hawksley P.G.W. (1967). Combustion of Pulverised Coal. The British Coal Utilisation Research Association.

[26]

Field, M.A. (1970). Measurements of the Effect of Rank on Combustion Rates of Pulverized Coal. Combust. and Flame, 14, pp. 237-248.

[27]

Fletcher,

T.H.

(1989).

Time-Resolved

Temperature

Measurements

of

Individual Coal Particles During Devolatilization. Combust. Sci. and Tech., 63, pp. 89-105. [28]

Fristrom, R.M., and Westenberg, A.A. (1965). Flame Structure. McGraw-Hill.

[29]

Glarborg, P., Kee, R.J., Grcar, J.F., and Miller, J.A. (1992). PSR: A Fortran Program for Modeling Well-Stirred Reactors. Sandia National Laboratories, Sandia Report SAND86-8209.

[30]

Glassman, I. (1977). Combustion. Academic Press.

[31]

Gokhale, A.J., Vasudevan, T.V. and Mahalingam R. (1986). Parametric studies on devolatilization of a subbituminous coal in a reactive gas environment. Fuel, 65, pp. 1670-1676.

[32]

Goldberg, P.M., and Essenhigh, R.H. (1978). Coal Combustion in a Jet-Mix Stirred Reactor. 17th Symp. (Int.) on Combust.. The Combust. Inst., PA, pp.145-154.

[33]

Gray, W.A., Kilham, J.K., and Muller, R. (1976). Heat Transfer from Flames. Paul Elek (Scientific Books) Ltd, London.

[34]

Hautman, D.J., Dryer, F.L., Schug, K.P., and Glassman, I. (1981). A Multiplestep Overall Kinetic Mechanism for the Oxidation of Hydrocarbons. Combust. Sci. Technol., 25, pp. 219-235.

49

[35]

Hertzberg, M., and Zlochower I.A. (1991). Devolatilization Wave Structures and Temperatures for the Pyrolysis of Polymethylmethacrylate, Ammonium Perchlorate, and Coal at Combustion Level Heat Fluxes. Combust. and Flame, 84, pp 15-37.

[36]

Hottel, H.C., Williams, G.C., Nerheim, N.M., and Schneider, G.R. (1965). Kinetic Studies in Stirred Reactors: Combustion of Carbon Monoxide and Propane. 10th Symp. (Int.) on Combust., The Combust. Inst., PA, pp. 111121.

[37]

Howard, J.B., Williams, G.C., and Fine, D.H. (1973). Kinetics of Carbon Monoxide Oxidation in Postflame gases. 14th Symp. (Int.) on Combust., The Combust. Instit., PA, pp. 975-986.

[38]

Howard, J.B. (1981). Fundamental of Coal Pyrolysis and Hydropyrolysis. In Chemistry of Coal Utilization. Second Supllementary Volume, Edited by Elliott M.A., John Wiley & sons, Inc., pp. 665-784.

[39]

Jachimowski, C.J. (1974). Kinetics of Oxygen Atom Formation During the Oxidation of Methane Behind Shock Waves. Combust. and Flame, 23, pp. 233-248.

[40]

Jensen, P.T., and Mitchell, R.E. (1993). High Temperature Char Reactivity Measurements in the Sandia Laminar Flow Reactor. Part 1: Text. Energy Research Project No. 1323/87-16. Lab. of Heating and Air Conditioning, Technical University of Denmark. Ministry of the Environment Geological Survey of Denmark.

[41]

Jones, W.P. and Lindstedt, R.P. (1988). Global Reaction Schemes for Hydrocarbon Combustion. Combust. and Flame, 73, pp. 233-249.

[42]

Kee, R.J., Grcar, J.F., Smooke, M.D., and Miller, J.A. (1985). A Fortran Program for Modeling Steady Laminar One-Dimensional Premixed Flames. Sandia National Laboratories, Sandia Report SAND85-8240.

[43]

Kee, R.J., Rupley, F.M., and Miller, J.A. (1992). Chemkin-II: A Fortran Chemical Kinetics Package for the Analysis of Gas Phase Chemical Kinetics. Sandia National Laboratories, Sandia Report SAND89-8009B (Reprinted).

[44]

Kobayashi, H., Howard, J.B., and Sarofim, A.F. (1976).Coal Devolatilization at High Temperatures. 16th Symp. (Int.) on Combust. The Combust. Inst., PA, pp. 411-425.

[45]

Kondratiev, V.N. (1958). Reactions of Some Radicals and Their Concentration in Flames. 7th Symp. (Int.) on Combust., The Combust. Inst., PA, pp. 41-44.

[46]

Kozlov, G.I. (1958). On High-Temperature Oxidation of Methane. 7th Symp. (Int.) on Combust., The Combust. Inst., PA, pp. 142-149.

[47]

LaNauze, R.D. (1982). Coal devolatilization in fluidized-bed combustors. Fuel, 61, pp. 771-774.

[48]

Laurendeau, N.M. (1978). Heterogeous Kinetics of Coal Char Gasification and Combustion. Prog. Energy Combust. Sci., 4, pp. 221-270.

[49]

Lavrov, N.V., and Karbirnichi-Kuznetsuv, V.B (1968). Nank. Uzb SSR, 25, pp. 9-.

[50]

Lester, T. W., Seeker, W.R., and Merklin, J. F. (1981). The influence of

50

oxygen and total pressure on the surface oxidation rate of bituminous coal. 18th Symp. (Int.) on Combust., The Combust. Inst., PA, pp. 1257-1265. [51]

Loison, R, and Chauvin, F. (1964). Pyrolyse Rapide du Charbon. Chimie et Industrie (Paris), 91, pp. 269-275.

[52]

Lyon, R.K., Hardy, J.E., and Von, Holt W. (1985). Combust. Flame, 61, pp. 79-.

[53]

Maloney, D.J., and Jenkins, R.G. (1984). Coupled Heat and Mass Transport and Chemical Kinetic rate Limitations During Coal Rapid Pyrolysis. 20th Symp. (Int.) on Combust., The Combust. Inst., pp. 1435-1443.

[54]

Miller, J.A., Mitchell, R.E., Smooke, M.D., and Kee,R.J. (1982). Toward A Comprhensive Chemical Kinetic Mechanism for the Oxidation of Acetylene: Comparison of Model Predictions with Results from Flame and Shock Tube Experiments. 19th Symp. (Int.) on Combust., The Combust. Inst., PA, pp. 181-196.

[55]

Miller, J.A., Smooke, M.D., Green, R.M., and Kee, R.J. (1983). Kinetic Modeling of the Oxidation of Ammonia in Flames. Combust. Sci. Technol., 34, pp. 149-176.

[56]

Miller, J.A., Branch, M.C., Mclean, W.J., Chandler, D.W., Smooke, M.D., and Kee,R.J. (1984). The Conversion of HCN to NO and N2 in H2-O2-HCN-Ar Flames at low pressure. 20th Symp. (Int.) on Combust., The Combust. Inst., PA, pp. 673-684.

[57]

Mitchell, R.E., and Madsen, O.H. (1986). Experimentally Determined Overall Burning Rates of Pulverized-Coal Char in Specified O2 and CO2 Environments. 21st Symp. (Int.) on Combust., The Combust. Inst., pp. 173-181.

[58]

Mitchell, R.E. (1988). On The Products of The Heterogeneous Oxidation Reaction at The Surfaces of Burning Coal Char Particles. 22nd Symp. (Int.) on Combust., The Combust. Inst., pp. 69-78.

[59]

Mitchell, R.E., Kee, R.J., Glarborg, P., and Coltrin, M.E. (1990). The Effect of CO Conversion in The Boundary Layers Surrounding Pulverized-Coal Char Particles. 23rd Symp. (Int.) on Combust., The Combust. Inst., pp. 11691176.

[60]

Ottaway M. (1982). Use of thermogravimetry for proximate analysis of coals and cokes. Fuel, 61, pp. 713-716.

[61]

Paczko, G., Lefdal, P.M., and Peters, N. (1986). Reduced Reaction Schemes for Methane, Methanol and Propane Flames. 21st Symp. (Int.) on Combust., The Combust. Inst., PA, pp.739-748.

[62]

Parikh, R.S., and Mahalingam, R. (1987). Modeling and Experimental Studies on Devolatilization Yields from a Bituminous Coal. Ind. Eng. Chem. Res., 26, pp. 2378-2384.

[63]

Peck, R.E., and Pabian, D.J. (1995). Flat-Flame Burner Studies of PulverizedCoal Combustion. Experimental Results on Flame Structure. Final Report. Prepared for Lab. of Heating and Air Conditioning, Technical University of Denmark. Prepared by Dep. of Mechanical and Aerospace Enginnering, Arizona State University, USA.

51

[64]

Pitt, G.J. (1962). The Kinetics of the Evolution of Volatile Products from Coal. Fuel, 41, pp. 267-274.

[65]

Ranz, W., and Marshall, W. (1952). Chem. Eng. Prog., 48, pp. 141-.

[66]

Rossberg,

M.

(1956).

Experimental

Results

Concerning

The

Primary

Reactions in the Combustion of Carbon. Z. Elektrochem., 60, pp. 952-. [67]

Saxena, S.C. (1990). Devolatilization and Combustion Characteristics of Coal Particles. Prog. Energy Combust. Sci., 16, pp. 55-94.

[68]

Seery, D.J., and Bowman C.T. (1970). An Experimental and Analytical Study of Methane Oxidation Behind Shock Waves. Combust. Flame, 14, pp. 37-47.

[69]

Serio, M.A., Hamblen, D.G., Markham, J.R., and Solomon, P.R. (1987). Kinetics of Volatile Product Evolution in Coal Pyrolysis: Experiment and Theory. Energy & Fuels, 1, No. 2, pp. 138-152.

[70]

Seshadri, K., and Peters, N. (1990). The Inert Structure of Methane-Air Flame. Combust. Flame, 81, pp. 96-118.

[71]

Siddal, R.G. (1974). Paper 16. Flux methods for the analysis of radiant heat transfer. J. of the Inst. of fuel., 47, pp. 101-109.

[72]

Singh, T., and Sawyer, R.F. (1971). CO Reactions in the Afterflame Region of Ethylene/Oxygen and Ethane/Oxygen Flames. Combust. and Flame, 13, pp. 403-417.

[73]

Slack, M.W. (1977) Rate Coefficient for H + O2 + M = HO2 + M Evaluated from Shock Tube Measurements of Induction Times. Combust. and Flame, 28, pp. 241-249.

[74]

Smith, I.W. (1971a). Kinetics of Combustion of Size-Graded Pulverized Fuels o

in the Temperature Range 1200-2270 K. Combust. and Flame, 17, pp. 303314. [75]

Smith, I.W. (1971b). The Kinetics of Combustion of Pulverized Semio

Anthracite in the Temperature Range 1400-2200 K. Combust. and Flame, 17, pp. 421-428. [76]

Smith, P.J. and Smoot, L.D. (1980). One-Dimensional Model for Pulverized Coal Combustion and Gasification. Combust. Sci. and Tech., Vol.23, pp. 1731.

[77]

Smoot, L.D., Horton, M.D. and Williams G.A. (1976a). Propagation of Laminar Pulverized Coal-Air Flames. 16th Symp. (Int.) on Combust. The Combust. Inst., PA, pp. 375-387.

[78]

Smoot, L.D., Hecker, W.C., and Williams, G.A. (1976b). Prediction of Propagating Methane-Air Flames. Combust. and Flame, 26, pp. 323-342.

[79]

Smoot, L.D. (1984). Modeling of Coal-Combustion Processes. Prog. Energy Combust. Sci., 10, pp. 229-272.

[80]

Smoot, L.D. (1991). Coal and Char Combustion. Fossil Fuel Combustion - A Source Book. Edited by Bartok W. & Sarofilm A.F., John Wiley & Sons, Inc., pp. 653-781.

[81]

Sobolev, G.K. (1959). High-Temperature Oxidation and Burning of Carbon Monoxide. 7th Symp. (Int.) on Combust., The Combust. Inst., PA, pp. 386391.

52

[82]

Solomon, P.R., Hamblen, D.G., Carangelo R.M., and Krause J.L. (1982). Coal Thermal Decomposition in An Entrained Flow Reactor: Experiments and Theory. 19th Symp. (Int.) on Combust.. The Combust. Inst., PA, pp. 11391149.

[83]

Solomon,

P.R.,

Hamblen,

D.G.,

Carangelo,

R.M.,

Serio,

M.O.,

and

Deshpande, G.V. (1988). General Model of Coal Devolatilization. Energy & Fuels, 2, No. 4, pp. 405-422. [84]

Solomon, R.P., Serio, M.A., and Suuberg, E.M. (1992). Coal Pyrolysis: Experiments, Kinetic Rates and Mechanisms. Prog. Energy Combust. Sci., 18, pp. 133-220.

[85]

Spalding D.B. (1963). The Art of Partial Modeling. 9th Symp. (Int.) on Combust., The Combust. Inst., PA, pp. 833-843.

[86]

Suuberg, E.M. (1977). Rapid Pyrolysis and Hydropyrolysis of Coal. Sc.D. thesis, Dep. of Chemical Engineering, Massachusetts Institute of Technology, Cambridge.

[87]

Tsatsaronis, G. (1978). Prediction of Propagating Laminar Flames in Methane, Oxygen, Nitrogen Mixtures. Combust. and Flame, 33, pp. 217-239.

[88]

Vandooren, J., Peeters, J., and Van Tiggelen, P.J. (1975). Rate Constant of The Elementary Reaction of Carbon monoxide with Hydroxyl Radical. 15th Symp. (Int.) on Combust., The Combust. Inst., PA, pp. 745-753.

[89]

Warnatz, J. (1984). Rate Coefficients in the C/H/O System. In Combustion Chemistry. Edited by Gardiner, Jr.. Springer-Verlag. pp. 197-360.

[90]

Water, B.J., Mitchell, R.E., Squire, R.G., and Laurendeau, N.M. (1988a). Overall Kinetic Parameters for Combustion of a Highly Non-Spherical Carbon Char. 22nd Symp. (Int.) on Combust., The Combust. Inst., PA, pp. 17-27.

[91]

Water, B.J., Squires, R.G., and Laurendeau, N.M. (1988b). Evidence for Formation of CO2 in the Vicinity of Burning Pulverized Carbon Particles. Combust. and Flame, 74, pp. 91-106.

[92]

Westbrook, C.K., and Dryer, F.L. (1979). A Comprehensive Mechanism for Methanol Oxidation. Combust. Sci. Technol., 20, pp. 125-140.

[93]

Westbrook, C.K., and Dryer, F.L. (1981a). Simplified Reaction Mechanisms for the Oxidation of Hydrocarbon Fuels in Flames. Combust. Sci. Technol., 27, pp. 31-43.

[94]

Westbrook, C.K., and Dryer, F.L. (1981b). Chemical Kinetics and Modeling of Combustion Processes. 18th Symp. (Int.) on Combust., The Combust. Inst., PA, pp. 749-767.

[95]

Westbrook, C.K., and Pitz, W.J. (1984). A Comprehensive Chemical Kinetic Reaction Mechanism for Oxidation and Pyrolysis of Propane and Propene. Combust. Sci. Technol., 37, pp. 117-152.

[96]

Westbrook, C.K., and Dryer, F.L. (1984). Chemical Kinetic Modeling of Hydrocarbon Combustion. Prog. Energy Combust. Sci., 10, pp. xxx-xxx.

[97]

Wicke, E., and Wurzbacher, G. (1962). Konzentrationsprofile Vor Einer Im Sauerstoffstrom

Verbrennenden

Kohlenstoffoberflache-I.

Ergebnisse. Int. J. Heat Mass Transfer, 5, pp. 277-289.

53

Experimentelle

[98]

Williams, G.C., Hottel, H.C., and Morgan, A.C. (1969). The Combustion of Methane in a Jet-Mixed Reactors. 12th Symp. (Int.) on Combust., The Combust. Inst., PA, pp. 913-925.

54

APPENDIX A1. Calculation of Composition of Volatiles from Raw Coal Knowledge of the mass fractions of single pyrolysis species (composition of volatiles or volatile components), e.g. YCO2, YH2O, YCO, YHCN, YC2H6, YCH4, YH2 and YH2S (YS), is very important. These mass fractions are usually unknown and must be calculated. The mass fractions for eight single pyrolysis species (YCO2, YH2O, YCO, YHCN, YC2H6, YCH4, YH2 and YS (YH2S)) can be calculated in following way Coal is anlysed in two ways: Ultimate and Proximate analysis. Ultimate analysis * C * VH * VO * VN * VS * V H2O * V Ash

V

Proximate analysis VM

(Volatile matter)

FC

(Fixed carbon)

* V H2O * V Ash

---------ΣVi = 1.00

---------ΣVi = 1.00 According to Solomon et al. (1992), there is no kinetic rate coefficient for sulfur and it equals the analysis value (the evolution of sulfur is not considered here).

A : V *8 = V *s B : ∑ V N,i =

(A1)

14 * 27 27 * * V 4 _ V 4 = V N,i = V N 14 14 27

(A2)

C : V *2 = V *H 2O D : ∑ V H,i =

(A3)

1 * 6 * 4 * * * V 4 + V 5 + V 6 +V 7 = V H 27 30 16

(A4)

1 * V4 27 === « V *6 = 6 Y *5 4 Y *7 + + 30 Y *6 16 Y *6 *

VH -

(A5)

55

E : ∑ V C,i =

=== «

12 * 12 * 12 * 24 * V1+ V 3+ V 4 + V 5 44 28 27 30 12 + V *6 + V *9 = V *C 16

(A6)

12 * 12 * ⎧ 12 * 24 * 12 * * *⎫ V 1 + V 3 = V C - ⎨ V 4 + V 5 + V 6 +V 9⎬ 28 30 16 44 ⎩ 27 ⎭

(A7)

F : ∑ V O,i =

32 * 16 * * V 1+ V 3=V O 34 28

(A8)

=== « V *1 =

34 ⎧ * 16 * ⎫ ⎨V O - V 3 ⎬ 32 ⎩ 28 ⎭

(A9)

*

*

*

*

Y1 and V3

V1

*

Y3

*

*

*

===> V3 = (V1 *Y3 )/Y1

*

Inserting V3 in Eq.(A9), gives

34 ⎧ * 16 Y *3 * ⎫ V 1⎬ ⎨V o 28 Y *1 ⎭ 32 ⎩

(A10)

⎧ 32 16 * ⎫ === « ⎨ + Y *3 ⎬ V *1 = V *0 ⎩ 34 28 Y 1 ⎭

(A11)

*

V 1=

and *

V 3=

28 ⎧ * 32 * ⎫ ⎨V 0 - V 1⎬ 16 ⎩ 34 ⎭

(A12)

*

Vi should be equal to VM (volatile matter). Corrected values of mass fraction of species can be calculated as * * * *H2O,corr = V *2,corr = V*2 VM V = = V CO2,corr V 1,corr V 1 n ∑V *i

(A14) (A13)

i=1

*

*

VM

*

V CO,corr = V 3,corr = V 3

(A15)

n

∑V

* i

i=1

*

*

*

V HCN,corr = V 4,corr = V 4

VM

(A16)

n

∑V

* i

i=1

56

*

*

VM

*

V C2H6,corr = V 5,corr = V 5

(A17)

n

∑V

* i

i=1

*

*

VM

*

V CH4,corr = V 6,corr = V 6

(A18)

n

∑V

* i

i=1

*

*

*

V H2,corr = V 7,corr = V 7

VM

(A19)

n

∑V

* i

i=1

*

*

*

V S,corr = V 8,corr = V 8

VM

(A20)

n

∑V

* i

i=1

⎞ ⎛ ⎟ ⎜ ⎜ * * * VM ⎟ = = ⎜V H2S,corr V 8,corr V 8 n * ⎟ Vi ⎟ ⎜ ∑ i=1 ⎠ ⎝

(A21)

where n

8

∑V = ∑V - V * i

i=1

* i

* H2O

(A22)

i=1

The mass fractions of volatiles used in the calculations are shown in Table A1. They are calculated from the adjusted values of Illinois #6 and Pittsburgh #8 coal (Table A2). Table A3 shows the analysis of raw coal. The values in column (2) are the adjusted values. In proximate analysis, the mass fractions of volatile matter, moisture and ash have the same values as the analysis values (column (1)); the mass fraction of fixed carbon is adjusted so that the total mass fraction is unity. In the ultimate analysis, the mass fractions of ash and moisture have the same values as the analysis values (column (1)); the other mass fractions, C, H, O, N, S, are adjusted so that the total mass fraction is also unity. In column (3) (only Illinois #6 coal) of the ultimate analysis table, the mass fractions of fixed carbon, ash and moisture are retained; the other mass fractions are adjusted. Table A4 shows the composition of volatiles (mass fraction of single species in volatile matter). These values are used to calculate composition of volatiles from unknown coal.

57

Table A5 shows the proximate and ultimate analysis of raw coal. Illinois #6

Pittsburgh #8

(a)

(b)

(c)

V

* CO2,corr

0.085

0.073

0.020

* V H2O,corr

0.066

0.073

0.032

V

* CO,corr

0.142

0.122

0.167

* V HCN,corr

0.024

0.028

0.029

* V C2H6,corr

0.082

0.090

0.168

* CH4,corr

0.045

0.049

0.055

V

* H2,corr

0.016

0.018

0.011

* V S,corr

0.046

0.053

0.016

Total

0.506

0.506

0.498

* H2O,corr

0.059

0.059

0.025

V

* vol

0.447

0.447

0.473

Total

0.506

0.506

0.498

V

V

*

Table A1: The adjusted mass fractions of volatiles. V vol is the adjusted volatile matter, equal to the analysis value times with the Q-factor (Q = 1.385 for Illinois #6 coal and Q = 1.50 for Pittsburgh #8 coal (Jensen and Mitchell (1993))). (a): The values used in Figs. D011, D012, D021 and D022. (b): The values used in the other calculations of Illinois #6 coal air flame. (c): The values used in all calculations of Pittsburgh #8 coal air flame. Illinois #6

Pittsburgh #8

(a)

(b)

(c)

0.309

0.307

0.430

*

0.203

0.189

0.233

* * b,act/V a,act

0.657

0.616

0.542

* * a,act/V vol

0.691

0.687

0.909

V

* * C2H6/V a,act

0.265

0.293

0.390

* * V C2H6/V vol

0.183

0.201

0.356

* a,act

* * CO+V HCN

V

=V

* * * C2H6+V CH4+V H2

+V

V b,=act = * * * V CO+V CH4+V H2 V V

Table A2: Ratio of mass fractions of active gases and C2H6 to mass fraction of total volatiles.

58

Proximate Analysis

Blair Athol (1)

(2)

(1)

(2)

(1)

(2)

(3)

(1)

(2)

(1)

(2)

(1)

(2)

% Volatile Matter (VM)

28.5

28.5

34.5

34.5

32.1

32.1

32.1

22.2

22.2

31.5

31.5

25.4

25.4

% Fixed Carbon (FC)

51.2

51.3

52.0

52.1

46.2

46.2

46.2

61.2

61.3

51.9

52.0

60.1

60.1

2.7

2.7

3.9

3.9

5.9

5.9

5.9

3.0

3.0

2.5

2.5

5.5

5.5

% Ash

17.5

17.5

9.5

9.5

15.8

15.8

15.8

13.5

13.5

14.0

14.0

9.0

9.0

Total

99.9

100.0

99.9

100.0

100.0

100.0

100.0

99.9

100.0

99.9

100.0

100.0

100.0

% Moisture

Ultimate Analysis

Blair Athol (1)

% Carbon % Hydrogen

(2)

65.2 64.81

Cerrejon

Illinois #6

Cerrejon (1)

(2)

Middleburgh

Illinois #6

Middleburgh

(1)

(2)

(3)

(1)

68.1 66.99

57.5

54.61

57.5

68.9

(2)

Pittsburgh #8

Pittsburgh #8 (1)

(2)

68.6

67.2

67.13

Ulan

Ulan (1)

(2)

68.3 67.69

4.2

4.18

4.87

4.79

4.21

4.00

4.00

3.77

3.75

4.66

4.65

3.97

3.93

% Nitrogen

1.62

1.61

1.47

1.45

1.20

1.14

1.20

1.72

1.71

1.39

1.39

1.67

1.66

% Sulfur

0.53

0.53

0.69

0.68

4.33

4.11

4.33

0.36

0.36

1.29

1.29

0.23

0.23

% Oxygen

8.72

8.67

12.9 12.69

15.2

14.44 11.27

9.12

9.08

9.05

9.04

12.1 11.99

2.7

2.7

3.9

3.9

5.9

5.9

5.9

3.0

3.0

2.5

2.5

5.5

5.5

17.5

17.5

9.5

9.5

15.8

15.8

15.8

13.5

13.5

14.0

14.0

9.0

9.0

% Moisture % Ash Total

100.47 100.0

101.43 100.0

104.1 4

100.0 100.0

59

100.3 7

100.0

100.0 100.0 9

100.77 100.0

Table A3: Analysis of raw coal. (1): as received; (2),(3): corrected values

60

Mass fraction *

Y1 = YCO2 * Y2

= YH2O

Zap North Dakota lignite

Gillette sub-bituminous

Montana Rosebud sub-bituminous

Illinois #6 bituminous

Pittsburgh #8 bituminous

0.100

0.099

0.100

0.074

0.011

*

*

*

*

(0.099) 0.094

(0.067) 0.062

0.102

(0.048) 0.044

(0.028) 0.022

Y3 = YCO

0.194

0.154

0.068

0.123

0.092

* Y4

= YHCN

0.018

0.022

0.020

0.026

0.031

* Y5

= YC2H6

0.095

0.158

0.127

0.081

0.190

* Y6

= YCH4

0.025

0.043

0.034

0.044

0.050

*

0.017

0.012

0.013

0.016

0.012

*

0.011

0.440

0.012

0.038

0.024

Y9 = YNH3

*

0.001

0.000

0.001

0.000

0.000

C (nonvolatile)

0.440

0.440

0.520

0.550

0.562

1.000

1.000

1.000

1.000

1.000

*

Y7 = YH2 Y8 = YS (organic) or * (Y8 = YH2S)

*

Σ Yi =

Table A4: The composition of volatiles (Solomon et al. (1992)).

*

: The values in the parentheses are the "adjusted" values.

61

Proximate Analysis

Blair Athol (1)

(2)

(1)

(2)

(1)

(2)

(3)

(1)

(2)

(1)

(2)

(1)

(2)

% Volatile Matter (VM)

28.5

28.5

34.5

34.5

32.1

32.1

32.1

22.2

22.2

31.5

31.5

25.4

25.4

% Fixed Carbon (FC)

51.2

51.3

52.0

52.1

46.2

46.2

46.2

61.2

61.3

51.9

52.0

60.1

60.1

2.7

2.7

3.9

3.9

5.9

5.9

5.9

3.0

3.0

2.5

2.5

5.5

5.5

% Ash

17.5

17.5

9.5

9.5

15.8

15.8

15.8

13.5

13.5

14.0

14.0

9.0

9.0

Total

99.9

100.0

99.9

100.0

100.0

100.0

100.0

99.9

100.0

99.9

100.0

100.0

100.0

% Moisture

Ultimate Analysis

Blair Athol (1)

% Carbon % Hydrogen

(2)

65.2 64.81

Cerrejon

Illinois #6

Cerrejon (1)

(2)

Middleburgh

Illinois #6

Middleburgh

(1)

(2)

(3)

(1)

68.1 66.99

57.5

54.61

57.5

68.9

(2)

Pittsburgh #8

Pittsburgh #8 (1)

(2)

68.6

67.2

67.13

Ulan

Ulan (1)

(2)

68.3 67.69

4.2

4.18

4.87

4.79

4.21

4.00

4.00

3.77

3.75

4.66

4.65

3.97

3.93

% Nitrogen

1.62

1.61

1.47

1.45

1.20

1.14

1.20

1.72

1.71

1.39

1.39

1.67

1.66

% Sulfur

0.53

0.53

0.69

0.68

4.33

4.11

4.33

0.36

0.36

1.29

1.29

0.23

0.23

% Oxygen

8.72

8.67

12.9 12.69

15.2

14.44 11.27

9.12

9.08

9.05

9.04

12.1 11.99

2.7

2.7

3.9

3.9

5.9

5.9

5.9

3.0

3.0

2.5

2.5

5.5

5.5

17.5

17.5

9.5

9.5

15.8

15.8

15.8

13.5

13.5

14.0

14.0

9.0

9.0

% Moisturey % Ash Total

100.47 100.0

101.43 100.0

104.1 4

100.0 100.0

62

100.3 7

100.0

100.0 100.0 9

100.77 100.0

Table A5: Analysis of raw coal. (1): as received; (2),(3): corrected values.

63

APPENDIX A2. Numerical Calculation of DAEM for the Nonisothermal Process ∞

*

tj

V i - V ij = ∫ exp{- ∫ k ij * Vi 0 0

E max

dt} f (E) dE = ∫

where

f i (E) =

i

tj

exp{- ∫ k ij

dt} f (E) dE i

(A23)

0

E min

⎧ ( - )⎫ 1 exp⎨- E i E2 0 ⎬ ‰ σ i (2π ) ⎩ 2σ i ⎭ Ei } k i = k 0i exp{RT

(A24) (A25)

For t2

E max

*

V i -V i 1 = ∫ exp{- ∫ k i 1 dt} f i (E) dE j = 1: * Vi t1 E min E max

*

{ {

} (E) dE ∫ 1 dt + ∫ 2dt + ∫ 3dt ] } (E) dE

t2

t3

V i -V i 2 = ∫ exp - [ ∫ k i 1 dt + ∫ k i 2dt j= 2: * Vi t1 t2 E min E max

*

j=3:

(A26)

t2

V i -V i 3 = ∫ exp - [ k i * Vi t1 E min

t3

]

fi

(A27)

t4

ki

t2

ki

fi

(A28)

t3

or *

E max

V i - V ij = ∫ exp{ - [ k i 1∆ t 1 + k i 2∆ t 2 + • • • • • + k ij+1 ∆ t j ] } f i (E) dE * Vi E min where ∆tj = tj+1 - tj

64

(A29)